Art Branched Polyesters

Embed Size (px)

Citation preview

  • 7/27/2019 Art Branched Polyesters

    1/33

    Branched polyesters: recent advances in synthesis and performance

    Matthew G. McKeeb, Serkan Unala, Garth L. Wilkesb, Timothy E. Longa,*

    aDepartment of Chemistry, Virginia Polytechnic Institute and State University, 124A Davidson Hall, Blacksburg, VA 24061, USA

    bDepartment of Chemical Engineering, Virginia Polytechnic Institute and State University, Blacksburg, VA 24061, USA

    Received 28 September 2004; revised 12 January 2005; accepted 12 January 2005

    Abstract

    The synthesis, characterization, physical properties, and applications of branched polyesters are discussed. This review

    describes recent efforts in the synthesis of statistically and tailored branched systems, and performance advantages compared to

    linear counterparts. In particular, an emphasis is placed on long-chain branching, where the branches are sufficiently long

    enough to form entanglements. Step-growth polymerization methodologies that employ various combinations of multi and

    mono-functional groups to achieve different levels of branching are reviewed in detail. The performance of branched

    polyesters, including behavior in dilute and semi-dilute solutions, and melt and solid-state properties are discussed. The

    implications of topological parameters including branch length, number of branches, and branching architecture on rheological

    performance are also reviewed. Although the majority of this review focuses on the synthesis and rheological behavior of

    branched polyesters, some discussion is devoted to the influence of branching on solid-state properties, sub-micron fiber

    formation, and controlled biodegradation for drug-delivery applications. Finally, a perspective of future directions in highperformance applications for branched polyesters is provided.

    q 2005 Elsevier Ltd. All rights reserved.

    Keywords: Branching; Polyesters; Rheology; Entanglements; Crystallization; Step-growth polymerization

    Contents

    1. Scientific rationale and perspective . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 508

    2. Synthesis of long-chain branched polyesters via step-growth polymerization . . . . . . . . . . . . . . . . . . . . . . . . . . . 509

    2.1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 509

    2.2. Synthesis of branched polyesters via A2 and B2 monomers in the presence of An or Bn (nO2) monomers 511

    2.3. Synthesis of branched polyesters via A2 and B2 monomers in the presence of An or Bn (nO2) Monomers

    and a monofunctional endcapping reagent . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 513

    2.4. Synthesis of branched polyesters via AB monomers in the presence of A2B monomers . . . . . . . . . . . . . . 514

    Prog. Polym. Sci. 30 (2005) 507539

    www.elsevier.com/locate/ppolysci

    0079-6700/$ - see front matter q 2005 Elsevier Ltd. All rights reserved.

    doi:10.1016/j.progpolymsci.2005.01.009

    * Corresponding author. Tel.: C1 540 231 2480; fax: C1 540 231 8517.

    E-mail address: [email protected] (T.E. Long).

    http://www.elsevier.com/locate/ppolyscihttp://www.elsevier.com/locate/ppolysci
  • 7/27/2019 Art Branched Polyesters

    2/33

    3. Characterization of branched polymers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 516

    3.1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 516

    3.2. Contraction factors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 516

    3.3. Endgroup analysis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 517

    4. Influence of branching on melt rheological properties: model systems and long-chain branched polyesters . . . . . 518

    4.1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 518

    4.2. Number of branches per chain and branch length . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 519

    4.2.1. Randomly branched polyesters . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 519

    4.2.2. Star-branched polyesters . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 522

    4.2.3. H-shaped and comb-branched polymers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 524

    4.3. Flow activation energy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 525

    5. The influence of branching on solution rheology properties in the semidilute regime . . . . . . . . . . . . . . . . . . . . . 526

    5.1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 526

    5.2. Effect of branching on the entanglement concentration . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 526

    5.3. Recent advances in electrospinning of long-chain branched polyesters . . . . . . . . . . . . . . . . . . . . . . . . . . 527

    6. Influence of branching on thermal properties of polyesters . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 529

    6.1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 529

    6.2. Glass transition . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 530

    6.3. Melting behavior and quiescent crystallization growth . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 531

    6.4. Controlling the biodegradation of aliphatic polyesters through branching . . . . . . . . . . . . . . . . . . . . . . . . . 532

    7. Conclusions and future directions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 534

    Acknowledgements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 534

    References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 539

    1. Scientific rationale and perspective

    Branched polymers are characterized by the

    presence of branch points or the presence of more

    than two end groups and comprise a class of polymers

    between linear polymers and polymer networks [1].

    Although undesirable branching can occur in many

    polymerization reactions, controlled branching is

    readily achieved. In fact, numerous studies on polymer

    structure-property relationships have shown that

    branched polymers display enhanced properties and

    performance for certain applications [2]. Long-chainbranched polymers offer significantly different physi-

    cal properties than linear polymers and polymer

    networks. For example, a low concentration of long

    chain branching in the polymer backbone influences

    melt rheology, mechanical behavior, and solution

    properties, while large degrees of branching readily

    affects crystallinity [3,4]. The strong influence of only

    one long chain branch per chain can be visualized by

    looking at Fig. 1. The slip-links along the polymer

    backbone represent entanglements with other chains.

    The linear polymer is free to diffuse along a tubeimposed by other chains, while it is obvious from

    Fig. 1b that the mobility of the long-chain branched

    polymer is restricted, and must diffuse through

    some other mechanism. Thus, it is not surprising that

    long-chain branched polymers exhibit very different

    properties where chain entanglements play a role.

    (a) (b)

    Fig. 1. Cartoon representing entangled linear chains (a), and long

    chain branched chains (b). The slip links represent entanglements

    due to other polymers.

    M.G. McKee et al. / Prog. Polym. Sci. 30 (2005) 507539508

  • 7/27/2019 Art Branched Polyesters

    3/33

    It is widely documented that a high degree of

    branching in a polymer backbone provides enhanced

    solubility, lower viscosity and lower crystallinity, for

    the case of symmetric chains that readily crystallize,than a linear polymer of equal molecular weight [5].

    Therefore, a fundamental understanding of branching

    and how it influences polymer properties is essential

    for tailoring a polymeric material for high perform-

    ance applications. Numerous types of branched

    polymers can be prepared using different polymeriz-

    ation techniques. In a living polymerization, multi-

    functional initiators or multifunctional linking agents

    yield well-defined star-branched polymers. Alkyl-

    lithium initiators are particularly efficient types of

    multifunctional initiators, and polyfunctional silyl

    halides are highly efficient multifunctional linking

    agents [2]. Comb polymers, which contain extensive

    branching along the polymer backbone, are syn-

    thesized in the presence of a polyfunctional coupling

    agent. Polyfunctional or multifunctional monomers of

    a functionality greater than two result in randomly

    branched polymers. Randomly branched polymers are

    often prepared by step-growth or chain polymeriz-

    ation in the presence of a multifunctional comonomer

    [1]. Highly branched (hyperbranched) polymers are

    prepared without gelation via the self condensation of

    an ABx monomer containing one A functionalgroup, and two or more B functional groups that

    are capable of co-reacting. Unlike dendrimers, which

    exhibit a regular, tree-like branch structure from a

    central core, hyperbranched polymers contain linear

    segments (defects) due to a more randomly branched

    architecture. Hyperbranched polymers are generally

    produced more easily than dendrimers and exhibit

    several similar properties [6,7].

    The effect of branching on polymers prepared by

    chain-growth polymerization and single site catalyzed

    polymerizations has received significant attention.However, structure/property relationships for

    branched polyesters are limited and further studies

    are needed [8]. Polyesters offer good mechanical and

    thermal properties and high chemical resistance at

    relatively low cost. Many polyesters, such as poly

    (ethylene terephthalate)s (PET), polycarbonates, bio-

    degradable aliphatic polyesters, and liquid crystalline

    polyesters are commercially available [9]. PET is

    utilized for a wide range of applications including

    injection-molding and blow-molding [10]. Processing

    of some polyesters, however, is limited due to

    insufficient melt strength and melt viscosity. For

    example, while aliphatic polyesters such as poly

    (butylene adipate) (PBA) and poly(butylene succi-nate) (PBS) decompose rapidly under natural environ-

    mental conditions and are replacing some commodity

    polymers due to environmental concerns, processing

    these resins is often difficult due to low melt strength

    and melt viscosity [1113]. Thus, many researchers

    have focused on modifying polyesters for enhanced

    melt strength and melt viscosity by introducing long

    chain branches into the polyester backbone. In this

    review, the synthetic methods for preparing various

    long-chain branched polyesters are reported. More-

    over, the influence of branching on polyester proper-

    ties for new high performance applications is

    discussed.

    2. Synthesis of long-chain branched polyesters

    via step-growth polymerization

    2.1. Introduction

    Multifunctional comonomer branching agents are

    introduced into polycondensation reactions to obtainlong-chain branched polyesters. Unlike short chain

    branches (SCB), a long chain branch (LCB) is long

    enough to entangle with other chains in the melt and

    concentrated solutions thereby drastically altering the

    flow properties. The critical molecular weight (Mc) is

    the minimum molecular weight at which a polymer

    chain entangles, as often measured by the molecular

    weight dependence of viscosity [14]. The value Mcseparates two regimes in the dependence of zero shear

    rate viscosity (h0) on weight average molecular

    weight (Mw

    ) for linear chains. Below Mc

    the value

    ofh0 scales directly with Mw and above Mc h0 scales

    with M3:4w . The value of Mc for a given polymer is

    directly related to the entanglement molecular weight

    (Me), which is typically determined from the plateau

    modulus G0N as shown in Eq (1),

    MeZrRT

    G0N(1)

    where r is the polymer melt density, R is the gas

    constant, and T is the absolute temperature. Fetters

    M.G. McKee et al. / Prog. Polym. Sci. 30 (2005) 507539 509

  • 7/27/2019 Art Branched Polyesters

    4/33

    et al. relatedMc andMe through the packing length (p),

    which is proportional to the cross-sectional area of a

    polymer chain [15]. Values ofMc were reported in the

    literature for several linear polyesters, including PET(3300 g/mol), poly(decamethylene succinate) (4600 g/

    mol), poly(decamethylene adipate) (4400 g/mol), and

    poly(decamethylene sebacate) (4500 g/mol) [16]. The

    parametersMc andMe are discussed further in Sections

    4.1 and 4.2 of this review.

    Hudson et al. showed that long-chain branching in

    the polymer backbone permits control over the

    rheology of the polymer [17]. More recent studies

    on the modification of polyesters with long-chain

    branching have involved the use of PET, an

    engineering thermoplastic, with good thermal and

    mechanical stability, high chemical resistance, and

    ease of processing [10,18]. Early work by Manaresi et

    al. describes the preparation of long-chain branched

    PET using low levels of trimesic acid [19]. Intrinsic

    viscosity measurements and the extent of reaction

    were reported along with the degree of branching.

    It is well known that polycondensation reactions

    with multifunctional comonomers may form an

    infinite molecular weight polymer network, or gel,

    above a certain multifunctional comonomer concen-

    tration or at high conversions. The onset of gelation

    occurs at a critical point of conversion during thepolymerization, and is dependent on the degree of

    functionality and the concentration of the multi-

    functional (fO2) branching agent. For example,

    polyester networks are prepared using dicarboxylic

    acids and tri- or tetrafunctional monomers [20]. The

    critical extent of reaction (ac) at which a polymer is

    predicted to form a gel is shown in Eq. (2).

    acZ1

    rCrpfK21=2(2)

    This equation is valid for polymerization mixtures

    with bifunctional A and B monomers and a multi-

    functional A monomer. In Eq. (2), r is the ratio of A

    functional groups to B functional groups and p is the

    ratio of A functional groups with fO2 to the total

    number of A groups. Low concentrations of multi-

    functional comonomers are used at low conversions to

    obtain long chain branching, and this method has

    yielded low molecular weight polymers. Neff et al.

    suggested the use of a monofunctional comonomer

    together with bifunctional and multifunctional mono-

    mers to overcome the gelation problem in high

    multifunctional comonomer concentrations or athigh conversions [21]. Manaresi et al. were first to

    report the preparation and characterization of

    PETs synthesized in the presence of a high content

    (O1 mol%) of trifunctional comonomer (trimethyl

    trimesate), as well as monofunctional comonomers

    (methyl 2-benzoylbenzoate) [22]. Rosu et al. reported

    branched PETs using multifunctional and monofunc-

    tional comonomers and subsequent solid-state pol-

    ymerization was employed to increase the molecular

    weight of the final product [23].

    Jayakannan and Ramakrishna synthesized high

    molecular weight branched PETs through the copo-

    lymerization of an A2 monomer with small amounts

    of an AB2 monomer [24]. However, as discussed in

    detail later in this review, insoluble crosslinked

    polymers were obtained at higher conversions.

    Hudson et al. synthesized and characterized branched

    PETs to study the balance between the branching

    reagents and endcapping reagents [17]. The objective

    of their study was to examine various branching

    agents used for PETs and other polyesters and their

    influence on polymer properties, both with and

    without an endcapping reagent. Molecular weightwas controlled via endcapping reagents on branched

    polyesters using a variety of branching agents.

    More recently, Yoon et al. studied the effects of

    multifunctional comonomers such as trimethylo-

    lethane (TME) and pentaerythritol on the properties

    of PET copolymers [18]. Molecular weights increased

    with increasing comonomer content while the mole-

    cular weight distribution broadened. Although solid-

    state mechanical properties did not differ significantly

    from linear analogues, the branched copolymers

    exhibited earlier shear-thinning onset in the meltcompared to linear PET. Moreover, the crystallization

    rates of the copolymers decreased with increasing

    comonomer content as would be expected. Similar to

    branched PETs, branched poly(butylene isophthalate)

    (PBI) and poly(butylene terephthalate) (PBT) were

    synthesized and characterized to investigate their melt

    and crystallization properties. Linear and branched

    PBIs were synthesized from dimethylisophthalate

    (DMI) and 1,4-butanediol (BD) in the presence of

    trifunctional comonomers [25]. Branched PBTs were

    M.G. McKee et al. / Prog. Polym. Sci. 30 (2005) 507539510

  • 7/27/2019 Art Branched Polyesters

    5/33

    synthesized by incorporating the trifunctional como-

    nomer, 1,3,5-tricarboxymethylbenzene [26].

    High molecular weight branched aliphatic polye-

    sters such as poly(ethylene succinate), poly(butylenesuccinate) (PBS), and poly(butylene adipate) (PBA),

    known as Bionollee polymers, were also prepared

    [27]. Bionollee polymers are used in a variety of

    applications, including film blowing, blow molding,

    extrusion coating and extrusion foaming [28]. Han

    et al. described synthetic conditions and thermal and

    mechanical properties for high molecular weight

    branched PBAs [11]. Ramakrishnan et al. reported

    the synthesis of a series of branched thermotropic

    liquid crystalline polyesters and their structural

    features [29]. Other novel branched polyesters

    such as branched poly(3-hydroxy-benzoates) and

    poly(4-ethyleneoxy benzoate) were synthesized by

    Kricheldorf et al. and Ramakrishnan et al., respecti-

    vely, [30,31].

    2.2. Synthesis of branched polyesters via A2 and B2monomers in the presence of An or Bn (nO2)

    monomers

    The most common method for synthesizing

    branched polyesters is via the addition of small

    amounts of tri- or tetrafunctional comonomers to thepolymerization. Manaresi et al. first reported the

    synthesis of branched PETs from dimethyl terephthal-

    ate (DMT) and ethylene glycol (EG) using a trifunc-

    tional branching agent, trimethyl trimesate (Fig. 2)

    [19]. To prevent gelation, only small amounts (!2%)

    of the trifunctional branching agent were used and the

    influence of long chain branching on PET properties

    was also reported. Although Manaresi et al. did not

    report the absolute molecular weights of the polymers,

    intrinsic viscosities in o-chlorophenol at 25 8C and the

    extents of reaction by end-group analysis were

    reported. After ester-interchange, only one type of

    functional group remained, i.e. the hydroxyl group of

    the hydroxy ethyl ester (Fig. 3). In the subsequent

    polycondensation step, high molecular weight PET is

    formed via the evolution of ethylene glycol. (Fig. 4)Weisskopf used different trifunctional agents to

    synthesize high molecular weight branched PETs

    [32]. Trimethylolpropane (TMP) was used as a

    trifunctional branching agent and pentaerythritol

    was a suitable tetrafunctional branching agent.

    Trimethylolethane (TME) and trimesic acid were

    also used as multifunctional comonomers (Fig. 5).

    Hess et al. recently described the syntheses of both

    linear and branched PETs [33]. Branched PETs were

    obtained via the ester-interchange route starting from

    DMT and a 2.5 M excess of EG. The reactions were

    performed in a stainless-steel reactor with different

    amounts (0.07 to 0.43 mol% with respect to DMT) of

    trimethylolpropane (TMP) (branching agent, Fig. 5a)

    present during the transesterification step. Transester-

    ification was catalyzed with the addition of manga-

    nese acetate at a maximum temperature of 230 8C.

    Following transesterification, polycondensation was

    catalyzed by antimony acetate at a maximum

    temperature of 290 8C under vacuum.

    Yoon et al. synthesized branched PETs in a similar

    manner with TME as a branching agent at concen-

    trations from 0.04 to 0.15 mol% [18]. Titanium

    isopropoxide was used as the catalyst for the

    polycondensation reaction. High molecular weight

    PET copolymers were obtained with broad molecular

    weight distributions. The thermal properties of the

    copolymers were not significantly influenced by the

    comonomers due to the low concentrations, however

    the branched PET displayed enhanced zero shear rate

    viscosity (h0) and shear thinning behavior. In a similar

    fashion, branched PBI and PBT samples were prepared

    O

    O O

    OHO

    OH

    OO

    O

    OO

    O

    (a) (b) (c)

    Fig. 2. Bifunctional and trifunctional monomers used in the

    synthesis of branched PET. (a) dimethyl terephthalate (DMT),

    (b) ethylene glycol (EG), and (c) trimethyl trimesate (trimethyl

    1,3,5-benzenetricarboxylate) (TMT).

    COOCH2CH

    2OH

    COOCH2CH

    2OH

    COOCH2CH

    2OHHOCH

    2CH

    2OOC

    COOCH2CH

    2OH

    Fig. 3. Hydroxy ethyl esters formed after the ester-interchange step

    during the polycondensation of PET.

    M.G. McKee et al. / Prog. Polym. Sci. 30 (2005) 507539 511

  • 7/27/2019 Art Branched Polyesters

    6/33

    using A2, B2, and A3/B3 type monomers. Branched

    PBIs were synthesized with the trifunctional branching

    agent tris(hydroxyethyl) isocyanurate (THEIC) during

    the polymerization reaction of dimethyl isophthalate

    (DMI) with 1,4-butanediol (BD) in the presence of aTi(OBu)4 catalyst (Fig. 6) [25]. The branched PBIs

    were prepared via a two step polycondensation

    reaction. In the first step, the reaction temperature

    was raised from 140 to 200 8C and held at 200 8C until

    about 90% of the theoretical amount of methanol was

    collected. In the second step, the pressure was reduced

    and the temperature was maintained in the range of

    200230 8C. The temperature was maintained lower

    than normally employed for polyesters, such as PBT, to

    prevent side reactions. Compositional and structural

    characterization included elemental analysis, mass

    spectroscopy, 1H NMR spectroscopy, HPLC, and end

    group analysis. Linear and branched PBTs were also

    recently synthesized using DMT and BD as bifunc-

    tional monomers and trimethyl trimesate (TMT) as a

    trifunctional comonomer [26]. Using titanium tetra-

    butoxide at 250 8C in the second step yielded randomly

    branched PBTs [28].

    Han et al. synthesized high molecular weight

    branched PBAs from aliphatic dicarboxylic acids and

    glycols in the presence of glycerol or pentaerythritol

    [11]. The influence of reaction parameters such as

    catalyst concentration, reaction time, temperature,

    and concentration of branching agent on molecular

    weight was examined. These branched PBAs were

    prepared via the synthesis of linear PBA from adipicacid and BD in the presence of a titanium(IV)

    isopropoxide (TIP) catalyst and a triethylamine

    (TEA) cocatalyst. The resulting linear polymer was

    reacted with adipic acid in the presence of TIP to

    obtain prepolymers with carboxylic acid end groups.

    In a second step, the carboxylic acid terminated PBA

    prepolymers were condensed with the branching

    agent (glycerol or pentaerythritol) in the presence of

    TIP to obtain branched PBS. Han et al. studied

    molecular weight with respect to multifunctional

    comonomer concentration and showed that both the

    molecular weight and the molecular weight distri-

    bution of branched PBAs increased with increasing

    concentration of glycerol up to 0.6 wt% relative to the

    PBA prepolymer. The gel content of the branched

    PBAs also increased with increasing glycerol con-

    centration up to 0.6 wt%. Surprisingly, 0.9 wt% or

    more glycerol resulted in lower gel content and lower

    molecular weights. The authors did not offer an

    explanation for this dependence of gel content and

    molecular weight on branching content.

    C C

    C

    O O

    OCH2CH2OOCH2CH2O CC

    O O

    CC

    O O

    O

    x y

    z

    OCH2CH2O C

    O

    C

    O

    Fig. 4. Structure of randomly branched PET.

    (c) (d)(b)(a)

    OOH

    O OH

    OHO

    HO OH

    OH

    OH

    HO

    HO

    OH

    OH

    HO OH

    Fig. 5. Trifunctional and tetrafunctional branching agents used in the synthesis of branched PET. (a) trimethylolpropane (TMP),

    (b) pentaerythritol, (c) trimethylolethane (TME), and (d) trimesic acid (TMA).

    M.G. McKee et al. / Prog. Polym. Sci. 30 (2005) 507539512

  • 7/27/2019 Art Branched Polyesters

    7/33

    When branched PBAs were prepared with either

    glycerol or pentaerythritol, it was found that the

    molecular weight of branched PBAs with pentaery-

    thritol was higher due to the higher degree of

    functionality. The introduction of a branching agent,

    TMP, to the polycondensation system of succinic acidand BD resulted in high molecular weight randomly

    branched poly(butylene succinate) (PBS) [34]. The

    esterification was conducted using a 1.0 to 1.1 ratio of

    succinic acid to BD under nitrogen in the presence of

    a titanium isopropoxide catalyst. The temperature was

    raised from 140 to 200 8C as water was removed. The

    ensuing polycondensation step was performed by

    introducing 0.10.5 wt% of TMP to the reaction

    mixture at 140 8C. The reaction temperature was

    raised to 240 8C and the reaction was completed at a

    pressure less than 1 Torr. Absolute molecular weightsand molecular weight distributions were determined

    using SEC (size exclusion chromatography) with a

    multi-angle laser light-scattering (MALLS) detector.

    The molecular weight distribution and the weight

    average molecular weight increased with increasing

    amounts of TMP, while the number-average molecu-

    lar weight decreased.

    It is possible to synthesize numerous types of

    branched polyesters by introducing A3/B3 or A4/B4monomers into a polymerization of A2 and B2

    monomers that involve transesterification and poly-condensation. Long chain branched polymers gener-

    ally have higher weight average molecular weights

    and broader molecular weight distributions compared

    with linear polymers synthesized at equivalent reac-

    tion conditions. In fact, a high level of multifunctional

    comonomer results in insoluble crosslinked systems

    or high gel content if the conversion proceeds too far.

    Long chain branching strongly influences thermal,

    mechanical, and rheological behaviors of polymers as

    discussed in subsequent sections.

    2.3. Synthesis of branched polyesters via A2 and B2monomers in the presence of An or Bn (nO2)

    monomers and a monofunctional endcapping reagent

    The introduction of long chain branches in

    polyesters is accomplished using low levels of a

    multifunctional comonomer and low conversions

    since gelation occurs at high levels of multifunctional

    comonomer and at higher conversions. However, the

    use of monofunctional comonomers in the presence of

    bifunctional and multifunctional monomers prevents

    or decreases gel content, and high molecular weight

    polymers with long chain branches are attainable at

    higher conversions. Neff et al. incorporated a mono-

    functional reagent as a chain terminator to prevent or

    decrease gel formation [21]. Branched PETs were

    synthesized from DMT, EG, and diethylene glycol

    (DEG) as bifunctional monomers, with trimellitic

    anhydride as the branching agent and stearic acid as a

    monofunctional reagent (Fig. 7). Manaresi et al.

    synthesized highly branched PETs using a two step

    polycondensation reaction in the presence of a

    monofunctional comonomer, methyl 2-benzoylbenzo-

    ate, with bifunctional monomers, DMT and EG, and

    the trifunctional monomer, trimethyl trimesate

    (Fig. 2) [22]. Munari et al. used a monofunctional

    comonomer to shift the gel point to higher percentconversions, and no gelation occurred when the ratio

    of monofunctional monomer to trifunctional mono-

    mer was greater than 3. When the ratio was less than

    3, the gel point was reached at lower conversions.

    Methyl 2-benzoylbenzoate was used as the mono-

    functional comonomer, and polycondensation tem-

    peratures caused an approximately 30 wt% loss of

    monofunctional comonomer.

    Rosu et al. recently reported the synthesis and

    characterization of high molecular weight branched

    PETs that were prepared using a two step polycon-densation reaction in the presence of monofunctional

    NC N

    CNC

    O

    O

    O

    OH

    OH

    HOO

    O

    O

    O

    HOOH

    (a) (b) (c)

    Fig. 6. Bifunctional and trifunctional monomers used in the

    synthesis of branched PBI. (a) dimethyl isophthalate (DMI),

    (b) 1,4-butanediol (BD), and (c) tris(hydroxyethyl) isocyanurate

    (THEIC).

    O

    O

    OO

    HOO

    HO

    (a) (b)

    Fig. 7. Trifunctional and monofunctional comonomers used by Neff

    et al.29 in the synthesis of branched PETs. (a) trimellitic anhydride,

    and (b) stearic acid.

    M.G. McKee et al. / Prog. Polym. Sci. 30 (2005) 507539 513

  • 7/27/2019 Art Branched Polyesters

    8/33

    dodecanol or benzyl alcohol, DMT and EG, and

    multifunctional monomers, glycerol or pentaerythritol

    [23]. The polymers with glycerol and pentaerythritol

    displayed different degrees of branching as pentaer-ythritol has four primary alcohol groups, while

    glycerol has two primary and one secondary alcohol

    group. Therefore, the degree of branching was

    expected to be higher with pentaerythritol. In addition

    to two-step polycondensation reactions, solid-state

    polymerization was used to obtain high molecular

    weights. Solid-state polymerization increases polye-

    ster molecular weight, while avoiding thermal degra-

    dation [35,36]. This method enables the preparation of

    linear and branched ultra-high-molecular-weight

    PETs with intrinsic viscosities of more than 2 dl/g

    (which corresponds to number-average molecular

    weights of 110,000 g/mole approximately) [37].

    Solid-state polymerization of PET is typically con-

    ducted 1535 8C below the melting point of the

    polymer for various times [35]. It is also possible to

    perform these reactions at various temperatures

    (220, 230, 235 8C) under vacuum [36]. Rosu et al.

    reported molecular weight control for branched PETs

    when the reaction time in the solid-state polymeriz-

    ation step was controlled and specific compositions of

    reagents were used [23]. The branched PETs were

    characterized by solution viscometry, thermal anal-ysis, and melt rheology.

    Recently, Hudson et al. studied branched PETs

    based on various branching agents and different

    endcapping reagent compositions [17]. Branched

    PETs were prepared using conventional polyconden-

    sation reactions with a monofunctional monomer and

    various multifunctional monomers with DMT and EG

    or bis(2-hydroxyethyl) terephthalate. In addition to

    branching agents such as glycerol, pentaerythritol,

    and benzene-1,3,5-tricarboxylic acid (trimesic acid),

    Hudson et al. used benzene-1,2,4,5-tetracarboxylicacid, dipentaerythritol, and tripentaerythritol as

    branching agents with the endcapping reagent benzyl

    alcohol (Fig. 8). A wide range of branched PETs (with

    various branching agents) as well as their compo-

    sitions with and without the presence of an end-capping reagent were subsequently reported. The

    polymers were characterized using FTIR and 1H

    NMR spectroscopy, light scattering, dilute solution

    viscometry, and melt rheology to investigate the

    influence of branching on solution and melt

    properties.

    Although end-group modification of linear PETs

    for enhanced solubility and blend compatibility was

    previously reported, Kim and Oh investigated the

    effect of functional end groups on the physical

    properties of PETs by synthesizing hydroxyl and

    carboxylic acid end-capped linear and branched PETs

    [38]. The end-capped polymers were characterized

    using NMR spectroscopy, viscosity measurements,

    SEC, and thermal analysis. The high molecular

    weight branched PETs (MwO100,000) had broad

    molecular weight distributions, and diethylene glycol

    (DEG) units were present in the polymer backbone,

    which was attributed to side reactions of ethylene

    glycol during polycondensation.

    2.4. Synthesis of branched polyesters via AB

    monomers in the presence of A2B monomers

    An alternate method for synthesizing branched

    polyesters involves the copolymerization of A2/B2 or

    AB monomers with AB2/A2B monomers Ramak-

    rishnan et al. reported the synthesis and character-

    ization of branched and kinked PETs through the

    copolymerization of an A2 monomer with small

    amounts of an AB2 monomer [24]. The term

    kinked describes linear disruption in the PET

    backbone due to meta substitution of the aromatic

    group. Therefore, in order to understand theinfluence of kinks, linear and branched polymers

    O

    O

    OH

    OH

    O

    O

    HO

    HO

    OH

    OH

    OHO OH

    OH

    OH

    OH

    OH

    OH

    OH

    OH

    OHO O

    OH

    OH

    (a) (b) (c)

    Fig. 8. Branching agents used by Hudson et al. (a) benzene-1,2,4,5-tetracarboxylic acid, (b) dipentaerythritol, and (c) tripentaerythritol.

    M.G. McKee et al. / Prog. Polym. Sci. 30 (2005) 507539514

  • 7/27/2019 Art Branched Polyesters

    9/33

    were also prepared. Branched and kinked PETs were

    synthesized using melt polymerization of bis-(2-

    hydroxyethyl) terephthalate (BHET) as A2 monomer

    and ethyl-3,5-(2-hydroxyethoxy)benzoate (EBHEB)as the AB2 monomer (Fig. 9). Linear and kinked

    PETs were synthesized via the polycondensation of a

    BHET monomer with a 3-(2-hydroxyethoxy) benzo-

    ate (E3HEB) monomer that has a 1,3 connectivity

    rather than a 1,4 connectivity, and the reaction was

    terminated early to prevent gel formation. Early gel

    formation was attributed to the fact that the EBHEB

    monomer behaved similarly to an A3 type instead of

    AB2, primarily due to the polycondensation reaction.

    Therefore, BHET was able to react with all three

    sites of EBHEB during polycondensation, whichresulted in gel formation during the early stages

    of the polymerization. Unfortunately, stopping

    the reaction early to avoid gelation yielded low

    molecular weight polymers.

    In addition to PETs, poly(4-ethyleneoxy benzoate)

    was synthesized using ethyl 4-(2-hyroxyethoxy)benzoate (E4HEB) as AB monomer and EBHEB as

    AB2 monomer [31]. Crosslinked polymers were

    formed at branching agent levels higher than

    50 mol%. When compared to branched PETs, the

    branching content in these materials was higher and

    therefore, a wider range of branched polymers were

    prepared to study the effect of branching on the

    thermal properties. Kricheldorf et al. prepared linear,

    long chain branched, and hyperbranched poly

    (3-hydroxy-benzoates) via condensation of acid

    chlorides, 3-(trimethylsiloxyl) benzoyl chloride asan AB type monomer, and 3,5-(bistrimethylsiloxyl)

    benzoyl chloride as AB2 type monomer [30].

    Fig. 9. Reaction schemes for the synthesis of branched and kinked PETs.

    M.G. McKee et al. / Prog. Polym. Sci. 30 (2005) 507539 515

  • 7/27/2019 Art Branched Polyesters

    10/33

    3. Characterization of branched polymers

    3.1. Introduction

    Since branching has such a dramatic influence on

    polymer properties, it is important to characterize

    polymer architecture on a molecular level. Short chain

    branches are recognized with spectroscopic methods,

    while sparsely long chain branched polymers are

    much more difficult to characterize. In practice, most

    branched polyesters are random in nature, with

    heterogeneous distributions of molecular weight,

    number of branch points, and length of the branches.

    Since random branching influences polymer molecu-

    lar weight and molecular weight distribution, it is

    important to deconvolute the effects of chain archi-tecture and molecular weight. The determination of

    molecular weight of a branched polymer using size

    exclusion chromatography (SEC) and a calibration

    curve based on linear polystyrene results in large

    errors since separation is based on hydrodynamic

    volume and linear and branched chains can possess

    the same hydrodynamic size, but different molecular

    weights [39]. Consequently, light scattering, a method

    that does not depend on any standards or shapes, is

    critical for measuring the absolute Mw of branched

    polymer chains [40].

    3.2. Contraction factors

    When compared in the same environment (tempe-

    rature and solvent), a branched polymer has a higher

    segment density and a lower hydrodynamic volume

    than that of a linear polymer of equal molecular

    weight. Solution or melt viscosity measurements,

    coupled with SEC and light scattering experiments

    yield information regarding polymer size [41]. The

    mean square radius of gyration, hR2gi, is a measure of a

    polymers hydrodynamic volume as measured using

    static light scattering. Consequently, the ratio of the

    hR2gi of a branched polymer to that of a linear polymer

    of the same molecular weight in the environment

    expresses the degree of branching, a quantity, g,

    referred to as the index of branching or contraction

    factor, shown in Eq. (3).

    gZhR2gibranched

    hR2gilinear(3)

    Similarly, the ratio of the intrinsic viscosity ([h]) of

    a branched chain to a linear chain, conventionally

    denoted as g 0, is employed as shown in Eq. (4).

    g0Z

    hbranchedhlinear

    (4)

    The value of g 0 is easily determined using the

    procedure of Hudson et al., where a multi-angle laser

    light scattering (MALLS) detector and a viscosity

    detector are coupled with SEC [17]. The value of

    [h]branched is measured directly using the viscosity

    detector and [h]linear is calculated using the Mark

    Houwink relationship shown in Eq. (5).

    hlinearZKMaw (5)

    The parameters, K and a, are the MarkHouwink

    constants for a linear polymer. For a linear chain, g

    and g 0 are equal to 1.0 and decrease as the level of

    branching increases. Fig. 10 shows the decrease ofg,

    denoted gM in the figure, for branched poly(vinyl

    acetate) as a function of molecular weight [42]. The

    contraction factor decreases from about 0.95 at low

    molecular weight to 0.45 at higher molecular weights,

    indicating the high molecular chains have a larger

    degree of branching.

    Since the value of [h] of a branched polymer is

    lower than that of its linear analog, the MarkHouwink exponent, a, for a branched polymer is

    generally smaller than that of a linear chain.

    Comparison of the intrinsic viscosity dependence on

    Mw for a series of linear and branched polystyrenes

    showed a systematic decrease in the MarkHouwink

    Fig. 10. Contraction factor vs. molecular weight for randomly

    branched poly(vinyl acetate). The contraction factor decreases from

    about 0.95 for low molecular weight species to 0.45 for the higher

    molecular weight chains, indicating the high molecular chains have

    a larger degree of branching.

    M.G. McKee et al. / Prog. Polym. Sci. 30 (2005) 507539516

  • 7/27/2019 Art Branched Polyesters

    11/33

    exponent from 0.73 to 0.68 to 0.39 for linear, star-

    branched, and hyperbranched topologies, respectively

    [43,44]. It should be noted that other researchers have

    showed that, in the limit of high molecular weights,linear polymer and star polymers that possessed

    a large range of arm numbers exhibited equal

    MarkHouwink exponents for polybutadienes and

    polyisoprenes [45,46]. This discrepancy between the

    dependence of [h] on molecular weight for star-

    shaped polymers may possibly be attributed to the

    lower polystyrene molecular weights that were

    investigated in Ref. [43]. Lusignan, Mourey, Wilson,

    and Colby utilized the disparity in the MarkHouwink

    exponent for linear and branched chains to estimate

    the average distance between branch points for a

    randomly branched polyester [47]. Fig. 11 shows the

    intrinsic viscosity as a function of Mw for SEC

    fractions of the branched polyester. The two slopes

    correspond to a values of 0.80 and 0.43, typical for

    linear and branched chains, respectively, and the two

    lines intersect at 66,000 g/mol. The authors concluded

    that this crossover from linear behavior to branched

    behavior marks the average linear chain length in the

    polymer system and the weight average molecular

    weight between branch points.

    The contraction factor has been theoretically

    correlated to the branching parameters of a polymer

    chain. Zimm and Stockmayer related g values to the

    average functionality and the number of branchingunits for randomly branched chains [48]. For star

    polymers with monodisperse arms under theta con-

    ditions, g can be calculated using Eq. (6),

    gZ3fK2

    f2(6)

    where fis the number of arms. Unfortunately, it is not

    possible to measure the mean square radius of

    gyration for low molecular weight chains due to the

    limitation of MALLS. In fact, hR2gi1=2 measurements

    are unreliable for values less than 10 nm, which isroughly a value ofMw on the order of 104 g/mol [49].

    Intrinsic viscosity and g 0 are more reliable measure-

    ments, however a theoretical basis is not developed

    that relates g 0 to molecular parameters since the

    dependence ofg 0 and g is not understood. Much work

    has focused on understanding this relationship, and

    empirical correlations suggest the form

    g0Zg

    3 (7)

    where 3 is between 0.5 and 1.5, and is dependent on

    the type of branching [50]. Generally, 3 is equal to 0.5

    for low levels of branching or for star polymers, while

    3 is closer to 1.5 for comb-shaped polymers [51].

    Jackson, Chen and Mays determined values of 3

    between 0.8 and 1.0 for randomly branched poly

    (methyl methacrylate), and concluded that the visco-

    metric radius (Rv) is more sensitive to the radius of

    gyration due to the higher segment density of the

    branched chains [52]. Instead of trying to convert

    intrinsic viscosity contraction factors to radius of

    gyration contraction factors, Balke et al. developed an

    empirical relationship between g 0 and number of

    arms, ranging for star-branched poly(methyl metha-crylate) with 3270 arms [53]. The authors showed

    the number of arms, f, was more accurately predicted

    through the empirical fits than by estimating values of

    3 and using Eqs. (6) and (7).

    3.3. Endgroup analysis

    The average number of branches per chain for a

    step-growth polymer is determined from the basic

    Fig. 11. Intrinsic viscosity as a function of Mw for a randomly

    branched polyester. The intersection of the two slopes correspond-

    ing to respective a values of 0.80 and 0.43 mark the average

    molecular weight between branch points.

    M.G. McKee et al. / Prog. Polym. Sci. 30 (2005) 507539 517

  • 7/27/2019 Art Branched Polyesters

    12/33

    theoretical concepts of Flory and Stockmayer. The

    models are dependent upon the number of polymer

    chain ends, molecular weights, and initial concen-

    tration of the mono-, bi-, and tri-functional repeatunits. End-group analysis on the branched polymer

    reveals the number of total end-groups. Researchers

    calculated the number of end groups for PET

    branched with a trifunctional compound from the

    concentration of hydroxyl and carboxylic acid end-

    groups [54]. Utilizing the number of end-groups,

    Manaresi et al. and others employed the number

    average branching density or the average number of

    branches per chain, Bn,

    BnZ

    2r

    3KrK3p(8)

    where r is a parameter that represents the initial

    polymer composition,

    rZ3ntrif

    3ntrifC2nbif(9)

    and ntrifis the number of initial trifunctional molecules,

    nbif is the initial number of bifunctional molecules,

    and p is the extent of reaction given by Eq. (10),

    pZ

    1K

    E

    Mb

    2K

    r

    Mb

    2K

    Mt

    3

    (10)

    where Eis the sum of all end-groups (equiv/g) and Mband Mt are the molecular weights of the bi- and tri-

    substituted repeat units, respectively.

    A mono-functional agent is often added to a step-

    growth reaction mixture in tandem with the multi-

    functional branching agent for facile control of

    molecular weight. Theory developed by Flory and

    Stockmayer predict that addition of a mono-functional

    agent shifts the conversion at which gelation occurs to

    higher values. Moreover, when the molar ratio of

    mono- to tri- substituted agents is greater than or equal

    to three, the gel point cannot be reached [22]. The

    addition of a mono-functional compound is easily

    accounted for in the above analysis by introducing the

    parameters,

    r0Z

    nmono

    nmonoC2nbifC3ntrif(11)

    where nmono is the number of moles of mono-

    functional agent. The extent of reaction, p, is

    redefined as,

    pZ1KEMb

    2C

    Mt

    3K

    Mb

    2

    rC MmK

    Mb

    2

    r

    0

    (12)

    where Mm is the molecular weight of the mono-

    functional unit. Finally, the average number of

    branches, Bn, is given by Eq. (13).

    BnZ2r

    3KrC3r0K3p(13)

    It should be noted that long chain branches are often

    below the detection limit of end-group analysis and

    dilute solution measurements for mixtures of linear

    and branched chains. Consequently, the aforemen-

    tioned methods are often insensitive to sparsely

    branched chains [55,56]. Typically spectroscopic

    techniques and SEC methods are limited to detection

    of branching levels of 1 branch point per 10,000

    carbons [57]. Moreover, in polymer systems that

    contain both short chain and long chain branches,

    like polyolefins, the above methods cannot discrimi-

    nate between the two thereby making LCB detection

    difficult. More recently 13C NMR measurements

    detected branching levels of 0.35 branches per

    10,000 carbons in polyethylenes [58]. Since the flow

    behavior of polymers is sensitive to long chainbranches at concentrations far below the detection

    limit of the above methods, rheology becomes the only

    feasible way to identify low levels of this type of

    branching.

    4. Influence of branching on melt rheological

    properties: model systems and long-chain

    branched polyesters

    4.1. Introduction

    The dependence of viscosity on shear rate for

    branched chains is very different from that of linear

    chains, aand varies with the chain architecture

    (random, star-branched, comb-branched, H-shaped,

    etc). Typically, long-chain branched polymers exhibit

    shear and extensional viscosities that are unobtainable

    with linear chains. For example at low shear rates,

    branched chains can exhibit a viscosity greater than

    100 times that of linear polymer of equal molecular

    M.G. McKee et al. / Prog. Polym. Sci. 30 (2005) 507539518

  • 7/27/2019 Art Branched Polyesters

    13/33

    weight, while at high shear rates, the branched

    polymer may exhibit a lower viscosity than the linear

    polymer due to enhanced shear thinning [59].

    The presence of less than one long-chain branchalong a polymer backbone on average is known to

    significantly alter the flow properties [60]. Only long-

    chain branches, branches with MwOMc, can greatly

    alter rheological properties, while short-chain

    branches (SCB) only do not affect the rheological

    behavior [61]. Several branching parameters influence

    rheological properties including branch length, degree

    of branching, and chain architecture (random, star-

    branched, comb, H-branched, etc.). Typically,

    branches are introduced in a random fashion during

    polymerization, thereby leading to a broad distri-

    bution of branch lengths and branch density. Conse-

    quently it becomes difficult to separate the effects of

    branching distribution, molecular weight distribution,

    and chain architecture.

    As stated earlier, PET and other linear polyesters of

    relatively low molecular weight and narrow molecular

    weight distribution display poor melt strength and

    shear sensitivity at typical processing conditions

    [62,63]. Additives, molecular weight and molecular

    weight distribution changes by chain extension during

    reaction or post reactor processing, branching, chain

    end functionalization, and controlled cross-linking areoften used to modify the melt rheology of polyesters

    [64,65]. In addition, long chain branched polymers

    display superior melt strength and extensional

    viscosity compared to their linear analogs, which

    aids in blow molding and other processing appli-

    cations [62]. This section describes the influence of

    branching on the melt and concentrated solution

    rheological properties of polyesters, with particular

    focus on the influence of the branching parameters

    including number of branches, branch length, and

    branch type on melt viscoelasticity.

    4.2. Number of branches per chain and branch length

    4.2.1. Randomly branched polyesters

    For a randomly branched polymer chain, the

    average number of branches per chain (degree of

    branching) and the branch length are coupled for a

    given molecular weigh t. For example, a

    100,000 g/mol polymer with an average of one branch

    per chain has an a ver age bra nc h l ength of

    33,300 g/mol from the relationship,

    MbZMw

    2BnC1

    (14)

    where Mb is the molecular weight between branch

    points, Mw is the total polymer molecular weight, and

    Bn is the average number of branches per chain. Thus,

    if a higher concentration of multi-functional agent

    was added to a step-growth polymerization to yield a

    polymer chain with the same overall molecular

    weight, but with three branches per chain, the average

    branch length becomes 14,300 g/mol. Since effects of

    branch number and branch length cannot be separated

    for equivalent molecular weights, this section

    describes the influences of both parameters on

    rheological properties.

    The dependence ofh0 on Mw is well established for

    linear, flexible chains. Two regimes are separated by a

    critical molecular weight (Mc), below which h0 scales

    directly with Mw and above which h0 generally scales

    with M3:4w . Chains with molecular weights below Mcare too small to entangle, while the high molecular

    weight chains are topologically constrained due to

    entanglement couplings. Researchers have shown a

    significant departure from the h0KMw relationship

    exists for branched chains due to the reducedhydrodynamic volume of the branched chains at low

    molecular weights, and increased entanglement coup-

    lings at higher molecular weights [66].

    Hess, Hirt, and Opperman varied the level of

    random branching in PET by adding different levels of

    trimethylolpropane (TMP) to the melt polymerization,

    and the branched PET possessed a lower h0 compared

    to linear chains of equal Mw (approximately

    50,000 g/mol) [33]. Fig. 12 shows the systematic

    decrease in the zero shear rate viscosity as the average

    number of branches per chain was increased from 0.1to 0.5. The parameter g*, which is the ratio of the zero

    shear rate viscosity of a branched and linear chain at

    equal molecular weight, also decreased with the

    average number of branches per chain. For a series

    of linear and randomly branched poly(butylene

    isophthalate) polymers, Munari et al. also showed a

    decrease in h0 with w0.5 branches per chain

    compared to a linear chain of equivalent Mw(55,000 g/mol) [67]. Moreover, when these same

    authors employed the correlation that relates the ratio

    M.G. McKee et al. / Prog. Polym. Sci. 30 (2005) 507539 519

  • 7/27/2019 Art Branched Polyesters

    14/33

    ofh0 of a branched and linear chain of equal Mw as

    developed by Ajroldi et al., the branching index

    increased from 0 to 1.0 and the zero shear rate

    viscosity systematically decreased by four orders of

    magnitude. Similar behavior was exhibited for

    branched PET and branched poly(butylene tereph-

    thalate) [54,68].

    The influence of the level of random branching on

    the rheological properties of aliphatic polyesters hasalso been investigated. Kim et al. observed a

    systematic decrease in shear thinning onset for

    branched poly(butylene succinate) (PBS) as the level

    of trifunctional branching agent was increased from 0

    to 0.5 wt% [34]. The onset of shear thinning behavior

    was attributed to a higher entanglement density of the

    branched chains [69]. Moreover, the authors observed

    an increase in h0 for branched PBS over linear PBS.

    However the authors of this review believe the

    enhancement in both shear thinning and h0 was more

    likely due to the significant increase in Mw and

    polydispersity (Mw/Mn) for the branched chains

    relative to the linear chains. Short-chain, ethyl and

    n-octyl branches were also introduced into poly

    (ethylene adipate) (PEA) and PBS. In contrast to

    long-chain branched systems, a decrease in melt

    viscosity and increase in shear thinning onset were

    observed when compared to linear PEA and PBS of

    approximately equal weight average molecular weight

    [70]. The ethyl and n-octyl branches were not long

    enough to form entanglements, and consequently shear

    thinning and h0 enhancement were not as pronounced.

    The melt viscosity of the branched polyester was lower

    than that of the linear polyester due to the reduced

    hydrodynamic volume of the branched chains. Leher-meier and Dorgan studied the influence of blending

    linear polylactide and polylactides that were randomly

    branched through a peroxide cross-linking reaction on

    the melt rheological properties [71]. The authors

    ensured that the polylactides did not undergo degra-

    dation at the rheological conditions by adding the

    stabilizer, tris(nonylphenyl) phosphite. They observed

    excellent control over the rheological performance

    with the blend composition. In particular, the authors

    reported an increase in h0 and decrease in the

    frequency at which shear thinning occurred with

    increasing blend compositions of the branched chain.

    Unfortunately, molecular weight information was not

    reported, and the branching structures of the polylac-

    tides were not characterized. Thus, it was difficult to

    assess the relationships between branch structure and

    rheological behavior.

    Lusignan, Mourey, Wilson, and Colby studied the

    linear viscoelastic properties of randomly branched

    polyesters with varying branch lengths [47]. The

    authors showed for low branch length of NZ2

    monomeric units, the chains were unentangled and

    accurately described by the Rouse model withouthydrodynamic or topological interactions [72]. More-

    over, further studies showed that entanglements

    between the randomly branched chains did not form

    for N!20, since the Rouse model was adequate for

    branched polyesters with up to 20 repeat units

    between branch points [73]. Branched polyesters

    with NZ900 were synthesized to demonstrate that

    topological constraints dominate the viscoelastic

    response of chains with branch lengths long enough

    to entangle [47]. The average molecular weight

    between branch points, Mb, was determined byanalyzing the intrinsic viscosity dependence of Mwas shown in Fig. 11. The Mb (66,000 g/mol) was

    defined as the crossover from linear to branched

    behavior, and was measured by the reduction of the

    [h] vs. Mw slope. Below Mb, h0 scaled with M3:6w

    which was consistent with experimental results for

    entangled linear chains, and above Mb, h0wM6:0w due

    to the increased entanglement constraints imposed by

    the branched chains. Fig. 13 shows that the Rouse

    model breaks down for N/NeO2, where N/Ne is

    Fig. 12. Systematic decrease in (h0,b/h0,l) as a function of average

    number of branches per chain for randomly branched PET withMwZ50,000 g/mol.

    M.G. McKee et al. / Prog. Polym. Sci. 30 (2005) 507539520

  • 7/27/2019 Art Branched Polyesters

    15/33

    the number of entanglements per branch, due to

    entangled dynamics as the viscosity exponent, s,

    varies significantly from the Rouse prediction of 1.33.

    Consequently, the authors concluded that Mbw2Mefor entanglements to dominate the flow behavior.

    Earlier experiments by Long, Berry, and Hobbs

    observed that the rheological behavior of randomly

    branched polymers is dependent on the branch length

    [74]. They observed a larger h0 for branched

    poly(vinyl acetate) compared to a linear chain of

    equal molecular weight when the average molecular

    weight between branches (Mb) was greater than

    28,000 g/mol. Using MeZ9,100 g/mol, which was

    reported by Fetters et al. for poly(vinyl acetate) [75],

    Mbw3Me for entanglements between branches to

    control the melt rheological performance. Several

    other researchers have also investigated the rheologi-

    cal response of randomly branched chains. Valles and

    Macosko showed that Mc is higher for randomly

    branched poly(dimethylsiloxane) (PDMS) chains

    compared to linear chains as measured by the Mwdependence of h0 [76]. Moreover, the number of

    branches per chain also influenced Mc, as Mc

    increased from 33,000 for the linear chain to 98,000

    and 110,000 g/mol for the trifunctional and tetrafunc-

    tional polymers, respectively. Unlike the randomly

    branched polyesters, the researchers observed aweaker h0KMw relationship for branched PDMS

    above Mc. However, when h0 was plotted against the

    product gMw (where g is the contraction factor) in Fig.

    14, a viscosity enhancement was observed for the

    randomly branched PDMS as seen previously with

    star-branched polyisoprene. Finally, Masuda et al.

    reported a dependence of viscoelastic properties on

    Mb for randomly branched polystyrene in 50 wt%

    solutions [77]. A clear plateau region was not

    Fig. 13. For N/NeO2 deviation from the Rouse models is evident

    due to entangled dynamics as the viscosity exponent, s, significantly

    varies from the Rouse prediction of 1.33. The solid line is the Rouse

    prediction for N/Ne!2, and a phenomenological equation that

    describes entanglements for N/NeO2.

    Fig. 14. Dependence of h0 on the product gMw for randomly

    branched PDMS. The triangles and squares correspond to tri and

    tetra functionally branched PDMS, respectively.

    M.G. McKee et al. / Prog. Polym. Sci. 30 (2005) 507539 521

  • 7/27/2019 Art Branched Polyesters

    16/33

    observed for Mb/Me!2, due to relaxation of the short

    backbone segments via Rouse-like motions, while

    branched chains with Mb/MeZ6 showed h0

    enhancement.

    4.2.2. Star-branched polyesters

    This review has focused on randomly branched

    polyesters that contain a distribution of branch

    lengths. Unlike randomly branched polymers, the

    synthesis of star polymers allows for a high degree of

    control over the molecular structure. Due to the well-

    defined chain architectures that result from more

    controlled synthetic strategies, star polymers have

    received much attention in the area of structure/

    property relationships. Fundamental investigations of

    the dynamics of star polymers will provide useful

    information for understanding the behavior of com-

    mercially produced, randomly branched materials

    [78]. At relatively low molecular weight, the viscosity

    of a star polymer is lower than its linear analog,

    however, the viscosity of the a star polymer increases

    faster with molecular weight and exceeds that of the

    linear analog at some specific molecular weight [79].

    This molecular weight dependence occurs because the

    star polymer exhibits a reduced hydrodynamic

    volume compared to a linear polymer due to the

    higher segment density, however, a competing effectarises since the star polymer possesses restricted chain

    motion due to the constraint that one end of the arm is

    anchored to the star core. Consequently, the branch

    point hinders reptation, and relaxation only occurs

    when the arm retracts back along the confining tube

    and seeks a new direction [80]. McLeish and Milner

    suggested two modes of relaxation of a star polymer.

    Short relaxations occur at the chain end, where the

    branch point does not restrict the arm, and long-scale

    relaxations occur near the star core [80]. Experimental

    results by Ye and Sridhar corroborated this theory andrelaxation times for a concentrated solution of

    polystyrene stars were 20 to 150 times greater than

    the relaxation times predicted for linear chains by

    reptation theory [81]. In addition, star polymer

    polymer blend miscibility was highly influenced due

    to the impenetrable core of the star from which the

    arms diffuse outward [82].

    As mentioned previously, the zero shear rate

    viscosity for linear polymers follows a power law

    dependence above the critical molecular weight for

    entanglements, Mc. However, for star-shaped poly-

    mers with Mw greater than Mc, the zero shear rate

    viscosity increases exponentially with the weight

    average arm molecular weight [83]. Fetters et al.observed that the value h0 of a star-branched chain

    does not depend on the total Mw, but only on the arm

    molecular weight, Ma. Thus, h0 is independent of the

    number of arms. Later, Fetters et al. showed that the

    h0 of a 3-arm polyisoprene star was approximately

    20% lower than that of a 4-arm star of equivalent Ma,

    while for fO4, the degree of functionality is saturated

    and the viscosity is only dependent on Ma [84]. These

    experimental results were consistent with previously

    developed theories that suggested 3-arm stars have an

    additional mechanism of stress relaxation that stars of

    higher functionality do not exhibit [85]. They

    proposed for 3-arm stars, an arm could relax if the

    branch point diffuses down one of the tubes. The

    rheology of stars with higher degrees of functionality

    was also studied. Pakula et al. performed linear

    viscoelastic studies on polybutadiene stars with a

    significantly larger number of arms than previously

    studied [86]. The authors observed a high frequency

    and low frequency relaxation corresponding to chain

    segmental motion and terminal response, respecti-

    vely. Fig. 15 shows stars with fZ64 and fZ128 arms

    display an additional transition in the terminal flowrange that is not present for stars with fZ4 arms. This

    additional transition for stars with high degrees of

    functionality is attributed to cooperative rearrange-

    ment of the colloidal or liquid like structures in the

    melt [87]. It should be noted this additional relaxation

    would not be applicable for randomly branched

    polymer chains.

    Since arm length controls the viscoelastic response

    for stars with a relatively few number of arms, it is

    important to understand the role of branch length in

    order to provide viscosity enhancement. Kraus andGruver observed for equal overall molecular weight, a

    3 arm star polymer with MaZ10Me showed viscosity

    enhancement over the linear analog [83]. However,

    rheological studies performed on a series of asym-

    metric poly(ethylene-alt-propylene) stars showed that

    the critical Ma for viscosity enhancement was less

    than 10Me. Gell et al. studied a series of 3-arm

    asymmetric stars where two branch lengths were kept

    constant and one was varied, thereby providing a

    constant molecular weight backbone [88]. The branch

    M.G. McKee et al. / Prog. Polym. Sci. 30 (2005) 507539522

  • 7/27/2019 Art Branched Polyesters

    17/33

    length ranged from Mb/MeZ0 to Mb/MeZ18, and was

    shorter than the backbone length (Mbb/MeZ38).

    Deviation from linear chain behavior and h0 enhance-

    ment were observed for Mb/Mew23, with consider-

    ably fewer entanglements per branch needed than for

    symmetric stars due to the nature of the very long

    backbone compared to the branch length. Dorgan et

    al. showed that viscosity enhancement for 4 and 6-arm

    star poly(lactic acid) occurred at approximately Mb/

    MeZ4 [89]. The authors observed that the viscosity

    enhancement factor, G,

    GZh0;bM

    h0:lM(15)

    increased more rapidly for the 6-arm versus the 4-arm

    star as shown in Fig. 16. In Eq. (15) h0,b and h0,l are

    the respective zero shear viscosities of branched

    and linear chains of equal molecular weight. This

    result is in disagreement with theoretical treatments

    and experimental results, which show viscosity to be

    only dependent on arm length, but independent of the

    number of arms [90]. The discrepancy was attributed

    to a combination of polydispersity, hydrogen bonding

    effects between ester groups, and the relatively short

    arm lengths of the star polymers.

    Claesson et al. prepared star-shaped polyesters

    composed of poly(3-caprolactone) (PCL) initiated

    from hydroxy-functional hyperbranched cores end-

    capped with methacrylate units [91]. The end groups

    served as a cross-linking agent for utilization in

    powder coating applications. Due to the narrow range

    of molecular weights studied, it was difficult to

    determine if the star polymer exhibited exponential

    or power law behavior, however, the h0 was an order

    of magnitude lower compared to a linear polyester of

    equal Mw. Since the Mw of the hyperbranched cores

    cannot be neglected in the total Mw, there was a

    dependence of arm number on the zero shear rate

    melt viscosity for the PCL stars. The PCL star

    polymers with methacrylate end groups were cured

    with ultraviolet (UV) light [92]. Gelation occurred

    within seconds of UV exposure based on the

    crossover point of the storage modulus (G 0) and the

    loss modulus (G 0). The time to reach gelation

    increased linearly with the molecular weight of the

    star polymer since the concentration of methacrylate

    end groups decreased.

    Fig. 15. Frequency dependence of polybutadiene stars with (a) fZ4,

    (b) fZ64 and (c) fZ128 arms. The vertical dashed lines correspond

    the frequencies associated with relaxation of the chain segment (us)

    and arm (uR), respectively.

    Fig. 16. Viscosity enhancement vs. branching length for 4 and 6 arm

    star polylactide.

    M.G. McKee et al. / Prog. Polym. Sci. 30 (2005) 507539 523

  • 7/27/2019 Art Branched Polyesters

    18/33

    4.2.3. H-shaped and comb-branched polymers

    H-shaped architectures are considered the simplest

    form of a comb polymer, where branching occurs only

    at the two ends of the backbone. Although, little workhas focused on the synthesis and rheological analysis

    of H-shaped polyesters, many structure/property

    studies have been performed on model H-shaped

    polymers. This section of the present review serves to

    summarize work performed on model polymer

    systems, as the results are applicable to polyester

    architectures. Roovers studied the melt rheology of

    H-shaped polystyrene and observed a decrease in h0for the branched polymers compared to linear analogs

    at low molecular weight, and an increase in h0 at

    high molecular weights similar to star polymers [93].

    Fig. 17 shows that the viscosity enhancement factor,

    G, increased faster for the H-shaped polymer than that

    of either 3 or 4 arm stars as a function of chain

    entanglement per branch (Mb/Me). This was attributed

    to an additional mode of relaxation for the H cross-bar

    which is not present for star polymers. Archer and

    Varshney corroborated this extra relaxation mode for

    H-shaped or multi-arm polybutadienes with three

    branches per chain end [94]. The authors observed a

    broader and lower frequency transition to the terminal

    region for the multi-arm polybutadiene compared to

    its linear counterpart, which was attributed to anincreased relaxation time of the cross-bar. Moreover,

    they showed the terminal relaxation time and h0enhancement were primarily controlled by the branch

    length when MbOMe and were relatively independent

    of the cross-bar molecular weight. Houli et al. also

    studied the rheological behavior of pom-pom type

    polymers with a much greater number of arms (fZ

    32)[95]. They also observed the dominant mechanism of

    terminal relaxation was arm relaxation. However,

    multi-arm polymers with Mb!Mc did not form

    entanglements, which was marked by power law

    behavior from the glass to the terminal region, typical

    of Rouse-like motions. This is similar to the

    unentangled behavior of high molecular weight

    hyperbranched polymers that relax via segments that

    are smaller than Me [96].

    Although model star and pom-pom polymers were

    extensively studied to determine the influence of

    branching on rheological properties, most commer-

    cially produced polymers are randomly branched.

    Interest in the rheological characterization of comb-

    branched chains may bridge the gap between the

    behavior of model star polymers and randomly

    branched commercial polymer. Noda et al. performed

    viscoelastic measurements with polystyrene combs

    and showed a lower h0 compared to linear polystyrene

    of equal Mw, however when compared at equivalent

    Rg, the combs displayed viscosity enhancement [97].

    Roovers and Graessley also performed rheological

    analyses on comb polystyrenes with backbonemolecular weights of 275,000 and 860,000 g/mol

    with approximately 30 branches per chain varying in

    Mw from 6,500 to 98,000 g/mol [98]. The comb

    polystyrenes showed a reduction in h0 when com-

    pared to linear chains of the same Mw and showed h0enhancement when compared at equivalent intrinsic

    viscosity. Surprisingly, this enhancement was not

    restricted to branch lengths above Me, and the h0enhancement was different for the combs with

    different molecular weight backbones. However, the

    authors found good agreement with the log GK

    Mb/Merelationship for stars when the comb molecular weight

    was normalized by the average end-to-end comb

    molecular weight (MEE/Me). Daniels et al. studied the

    linear rheological response of comb-branched poly-

    butadiene and varied the molecular weight of the

    polymer backbone, the molecular weight of the arms,

    and the number of arms [99]. The researchers reported

    the viscoelastic response was dependent on the

    number of arms for low frequencies. At short time

    scales, the comb polymers displayed Rouse-like

    Fig. 17. Viscosity enhancement factor, G, as a function of the

    number of entanglements per branch. Triangles and diamonds

    denote H-polymers, the solid line denotes a 4-arm star, and the

    dashed line denotes a 3-arm star.

    M.G. McKee et al. / Prog. Polym. Sci. 30 (2005) 507539524

  • 7/27/2019 Art Branched Polyesters

    19/33

    behavior similar to star polymers due to relaxation of

    the dangling chain ends. For a relatively small number

    of arms, the comb polymers behaved similar to H-

    shaped polymers, marked by relaxation of the arms atintermediate frequencies and reptative motion of the

    backbone at low frequencies [100]. For a larger

    number of arms, the terminal region showed a

    distinctly different response for the comb compared

    to the H-shaped polymer, as a larger number of

    relaxation modes were available to the comb-

    branched polybutadiene. Roovers and Toporowski

    attributed the broader low frequency relaxation to

    additional couplings between a branch and a back-

    bone that are unavailable to star polymers [101].

    Namba et al. showed that branch spacing in comb

    polymers is important in their viscoelastic response

    [102]. They observed that highly branched comb

    poly(methyl methacrylate), with approximately one

    polystyrene branch (MbZ3450) per repeat unit did not

    entangle, as a plateau region was not observed for the

    storage modulus. Although Mb!Mc for the poly-

    styrene branches, the backbone molecular weight was

    well above Mc, so the lack of entanglements was

    attributed to the high branch density of the comb

    polymer. When the polystyrene macromonomer was

    copolymerized with methyl methacrylate to yield a

    branch structure of approximately 3 branches per 100repeat units, a clear plateau region was observed in the

    dynamic shear modulus. However, the authors did not

    address the incompatibility issues between poly-

    styrene and poly(methyl methacrylate) chains.

    Tsukarhara et al. also observed fewer entanglement

    couplings for highly branched combs with a backbone

    molecular weight greater than Mc. The researchers

    calculated Me of a highly branched poly(methyl

    methacrylate) comb from the plateau region, where

    Mb!Mc, and discovered Me was approximately 3

    orders of magnitude larger for the highly branchedPMMA compared to linear PMMA [103]. The authors

    attributed this to the increased cross-sectional area of

    the highly branched PMMA comb, which excluded

    other chains from a unit volume and thereby hindered

    entanglement couplings.

    4.3. Flow activation energy

    The activation energy explains the temperature

    dependence of viscosity for as shown in Eq (16),

    h0TZA0expEa

    RT

    (16)

    where h0 is the zero shear rate viscosity, A0 is a pre-exponential factor, Ea is the activation energy, R is the

    gas constant, and Tis the temperature in K. Generally,

    Eq. (14) is only valid for temperatures ca. 80 8C above

    the polymer glass transition (Tg) due to the exponential

    rise in viscosity at temperatures near the Tg. The value

    of Ea is independent of molecular weight and is only

    dependent on the local segmental nature of the chain

    [104]. Typical values for the Ea are in the range of 5 to

    30 kcal/mol. In general, Ea increases with either chain

    stiffness or bumpiness [105]. Consequently, the

    temperature dependence of viscosity for branched

    polymers differs significantly from the corresponding

    linear analogs. In particular, the rheological behavior

    of the former shows a greater temperature dependence

    and thus Ea is enhanced. One of the most outstanding

    examples is that Ea depends on the degree of

    branching and branch length in polyethylene, as

    several researchers reported a larger Ea for low-

    density polyethylene compared to high-density poly-

    ethylene. However, limited work has focused on the

    influence of branching on Ea in polyesters.

    Munari et al. investigated the influence of long-

    chain branching on the flow activation energy for aseries of partially aromatic polyesters, and found

    inconsistent results for the different polyesters. They

    reported a 35 to 100% increase in Ea for a series of

    branched poly(butylene terephthalate)s (PBT) com-

    pared to their linear analogs, however, only a slight

    increase in the Ea for branched PET compared to

    linear PET was observed [54,68]. Moreover, no

    enhancement in Ea was observed for branched

    poly(butylene isophthalate) (PBI) compared to linear

    PBI [106]. The authors attributed these discrepancies

    to differences in Mb between the branched polyestersand different temperature coefficients of the repeat

    units. Graessley related this inconsistent behavior to

    differences in the temperature coefficients of linear

    and branched polymer melts [107]. This discrepancy

    arises when considering the mode by which entangled

    chains relax. In particular, linear chains undergo

    reptation, while long chain branches relax by retrac-

    tion or short time-scale fluctuations along the tube

    contour length of an arm. As an arm relaxes, it must

    pass through a higher energy barrier due to the more

    M.G. McKee et al. / Prog. Polym. Sci. 30 (2005) 507539 525

  • 7/27/2019 Art Branched Polyesters

    20/33

    compact conformational states that are dependent on

    the temperature coefficient of the chain. In the cases

    where differences in activation energy for linear and

    branched polymers are observed, the quantity, DEZ

    (Ea)BK(Ea)L, is often used to quantify the degree of

    Ea enhancement. The quantity DE was shown to

    exponentially increase with the number of entangle-

    ments per branch (Mb/Me), and decrease to zero with

    decreasing branch length [108].

    5. The influence of branching on solution rheology

    properties in the semidilute regime

    5.1. Introduction

    The dilute solution properties of branched chains

    are consistent with their reduced hydrodynamic

    dimensions compared to linear polymers of equivalent

    molecular weight. The properties of branched poly-

    mers in dilute solution were discussed in some detail

    in Section 3.2. In dilute solution, the polymer chains

    are widely separated from each other, and only the

    interactions between two chains need to be con-

    sidered. At a critical concentration, C*, the polymer

    chains begin to crowd each other and overlap in

    solution, and COC* is termed the semidilute regime.

    As the polymer chains overlap, intrachain interactionsare screened at length scales longer than the

    correlation length, where the correlation length is

    defined as the average distance between neighboring

    contacts points [109]. Polymer concentrations above

    C* do not indicate that entanglement couplings

    between chains have formed [110]. Consequently,

    in the semidilute unentangled regime, C*!C!Ce(where Ce is the entanglement concentration), chain

    overlap is not sufficient to topologically constrain the

    polymer chain motion. Above Ce, the semidilute

    entangled regime, chain crowding and interpenetra-tion is sufficient to constrain the chain motion, and

    topological interactions dominate at distances longer

    than the tube diameter [111]. Limited rheological

    studies have shown that typically Ce/C* is in the range

    of 510 for neutral, linear polymers [111]. Finally, as

    polymer concentration is increased further into the

    concentrated regime, which is defined as the point

    where chain dimensions become independent of

    concentration, the polymer coils are highly entangled

    and behave similar to a melt [112].

    5.2. Effect of branching on the entanglement

    concentration

    The entanglement concentration is experimentallymeasured by analyzing the concentration dependence

    of specific viscosity (hsp),

    hspZh0Khs

    hs(17)

    where h0 is the zero shear rate viscosity of the

    polymer solution and hs is the solvent viscosity. For

    neutral, linear polymers in a good solvent, hspwC1.0

    in the dilute regime, hspwC1.25 in the semidilute

    unentangled regime, hspwC3.8 in the semidilute

    entangled regime as predicted by the reptation theory.

    Finally, the value of hsp generally shows a weakerdependence in the concentrated regime compared to

    the semi-dilute entangled regime [111]. For example,

    Colby et al. measured the onset of the semidilute

    unentangled and semidilute entangled regime for an

    aqueous solution of sodium hyaluranote [113]. Fig. 18

    shows the concentration dependence of viscosity and

    the determination of C* and Ce from the change in

    slope. Takahashi et al. measured h0 for linear poly

    (a-methylstyrene) in good, poor, and q solvents and

    observed the transition to the semidilute regime

    decreased with molecular weight as expected sincelarger chains begin to overlap at lower concentrations

    compared to smaller chains [114]. Moreover, the

    authors reported that the transition was dependent on

    Fig. 18. Concentration dependence of specific viscosity for a

    biopolymer, sodium hyaluranote. C* and Ce were determined as

    0.59 and 2.4 mg/mL, respectively.

    M.G. McKee et al. / Prog. Polym. Sci. 30 (2005) 507539526

  • 7/27/2019 Art Branched Polyesters

    21/33

    solvent quality. Other researchers investigated the

    hspKC relationship for linear poly(a-methylstyrene)

    in solvents of variable quality [115]. The investigators

    discovered that in dilute solutions hsp was lower inpoor solvents compared to good solvents, while the

    opposite was true in entangled solutions. It was

    proposed that the viscosity increase in the poor

    solvent was attributed to enhanced entanglement

    couplings due to relatively weak interactions between

    the chain segments and the solvent. Moreover, other

    researchers have observed a weaker concentration

    dependence for viscosity in the semidilute entangled

    regime for a polymer in a good solvent compared to a

    polymer in a theta solvent [116].

    Our discussion of concentration dependence ofhspand the determination of C* and Ce has focused on

    linear chains to this point. Not surprisingly, at equal

    molecular weights, branched chains show different

    behavior in solution compared to linear chains.

    Generally, as shown in Fig. 19, a branched polymer

    exhibits a higher overlap concentration compared to a

    linear polymer of equal Mw since the branch points

    act as obstacles to chain interpenetration [117].

    Sendijarevic et al. studied the effect of branching of

    AB/AB2 etherimide copolymer solutions on solution

    rheology properties [118]. The authors observed a

    weaker concentration dependence of h0 in thesemidilute entangled regime for more highly branched

    structures. As the copolymer composition was varied

    from 0 to 100 mol% AB2, corresponding to linear and

    hyperbranched architectures, respectively, the expo-

    nent decreased from 12.5 to 4.7, attributed to a largernumber of entanglements per chain in the linear

    copolymers. Moreover, the overlap concentration

    increased with higher levels of branching. Similarly,

    solution rheology studies with linear and randomly

    branched poly(ethylene terephthalate-co-ethylene iso-

    phthalate) (PET-co-PEI) showed a weaker hsp vs. C

    dependence for branched copolyesters compared to

    linear copolyesters of similar molecular weight [119].

    Furthermore, branching dramatically influenced the

    onset of the entanglement regime, as Ce increased

    from 4.5 to 10 wt% as g 0 for PET-co-PEI decreased

    from 1.0 to 0.43. In a related study, Juliani and Archer

    studied the rheology of unentangled and entangled

    A3KAKA3 multi-arm polybutadiene solutions with

    equivalent branch molecular weights and variable

    backbone molecular weights [120]. The investigators

    reported Ce decreased from 22 to 8 vol% as the

    molecular weight of the cross-bar was increased by

    w30% allowing a larger number of entanglements per

    chain for the higher molecular weight crossbar.

    5.3. Recent advances in electrospinning of long-chain

    branched polyesters

    Traditional melt processing of polyester fibers has

    received significant commercial attention, however,

    electrospinning has recently emerged as a specialized

    processing technique for the formation of sub-micron,

    high surface area fibers [121]. The utility of branched

    polyesters in the electrospinning process has received

    only limited attention. Typically, conventional poly-

    mer fibers are melt spun using pressure-driven flow

    through an extruder, yielding fibers on the order of

    10100 m