Вы находитесь на странице: 1из 33

Prog. Polym. Sci. 30 (2005) 507539 www.elsevier.

com/locate/ppolysci

Branched polyesters: recent advances in synthesis and performance


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

Department of Chemistry, Virginia Polytechnic Institute and State University, 124A Davidson Hall, Blacksburg, VA 24061, USA b Department 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 sufciently 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 inuence of branching on solid-state properties, sub-micron ber formation, and controlled biodegradation for drug-delivery applications. Finally, a perspective of future directions in high performance applications for branched polyesters is provided. q 2005 Elsevier Ltd. All rights reserved.
Keywords: Branching; Polyesters; Rheology; Entanglements; Crystallization; Step-growth polymerization

Contents 1. 2. Scientic rationale and perspective . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 508 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. 2.4. Synthesis of branched polyesters via A2 and B2 monomers in the presence of An or Bn (nO2) Monomers and a monofunctional endcapping reagent . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 513 Synthesis of branched polyesters via AB monomers in the presence of A2B monomers . . . . . . . . . . . . . . 514

* Corresponding author. Tel.: C1 540 231 2480; fax: C1 540 231 8517. E-mail address: telong@vt.edu (T.E. Long). 0079-6700/$ - see front matter q 2005 Elsevier Ltd. All rights reserved. doi:10.1016/j.progpolymsci.2005.01.009

508

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

3.

Characterization of branched polymers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3.1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3.2. Contraction factors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3.3. Endgroup analysis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Inuence of branching on melt rheological properties: model systems and long-chain branched polyesters . . . . . 4.1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4.2. Number of branches per chain and branch length . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4.2.1. Randomly branched polyesters . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4.2.2. Star-branched polyesters . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4.2.3. H-shaped and comb-branched polymers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4.3. Flow activation energy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . The inuence of branching on solution rheology properties in the semidilute regime . . . . . . . . . . . . . . . . . . . . . 5.1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5.2. Effect of branching on the entanglement concentration . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5.3. Recent advances in electrospinning of long-chain branched polyesters . . . . . . . . . . . . . . . . . . . . . . . . . . Inuence of branching on thermal properties of polyesters . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6.1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6.2. Glass transition . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6.3. Melting behavior and quiescent crystallization growth . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6.4. Controlling the biodegradation of aliphatic polyesters through branching . . . . . . . . . . . . . . . . . . . . . . . . .

516 516 516 517 518 518 519 519 522 524 525 526 526 526 527 529 529 530 531 532

4.

5.

6.

7.

Conclusions and future directions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 534 Acknowledgements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 534 References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 539

1. Scientic 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-chain branched polymers offer signicantly different physical properties than linear polymers and polymer networks. For example, a low concentration of long chain branching in the polymer backbone inuences melt rheology, mechanical behavior, and solution properties, while large degrees of branching readily affects crystallinity [3,4]. The strong inuence 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 tube imposed 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) 507539

509

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 inuences polymer properties is essential for tailoring a polymeric material for high performance applications. Numerous types of branched polymers can be prepared using different polymerization techniques. In a living polymerization, multifunctional initiators or multifunctional linking agents yield well-dened star-branched polymers. Alkyllithium initiators are particularly efcient types of multifunctional initiators, and polyfunctional silyl halides are highly efcient multifunctional linking agents [2]. Comb polymers, which contain extensive branching along the polymer backbone, are synthesized 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 polymerization 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 functional group, 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 signicant 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, biodegradable 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 insufcient melt strength and melt viscosity. For example, while aliphatic polyesters such as poly (butylene adipate) (PBA) and poly(butylene succinate) (PBS) decompose rapidly under natural environmental conditions and are replacing some commodity polymers due to environmental concerns, processing these resins is often difcult 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. Moreover, the inuence of branching on polyester properties 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 obtain long-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 ow 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 Mc separates two regimes in the dependence of zero shear rate viscosity (h0) on weight average molecular weight (Mw) for linear chains. Below Mc the value of h0 scales directly with Mw and above Mc h0 scales 3:4 with Mw . 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 G0 N as shown in Eq (1), Me Z rRT G0 N (1)

where r is the polymer melt density, R is the gas constant, and T is the absolute temperature. Fetters

510

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

et al. related Mc and Me through the packing length (p), which is proportional to the cross-sectional area of a polymer chain [15]. Values of Mc 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 parameters Mc and Me 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 modication 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 innite molecular weight polymer network, or gel, above a certain multifunctional comonomer concentration or at high conversions. The onset of gelation occurs at a critical point of conversion during the polymerization, and is dependent on the degree of functionality and the concentration of the multifunctional (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). 1 r C rpf K 21=2

ac Z

(2)

This equation is valid for polymerization mixtures with bifunctional A and B monomers and a multifunctional 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 multifunctional 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 monomers to overcome the gelation problem in high multifunctional comonomer concentrations or at high conversions [21]. Manaresi et al. were rst 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 monofunctional comonomers and subsequent solid-state polymerization was employed to increase the molecular weight of the nal product [23]. Jayakannan and Ramakrishna synthesized high molecular weight branched PETs through the copolymerization 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 inuence on polymer properties, both with and without an endcapping reagent. Molecular weight was 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 trimethylolethane (TME) and pentaerythritol on the properties of PET copolymers [18]. Molecular weights increased with increasing comonomer content while the molecular weight distribution broadened. Although solidstate mechanical properties did not differ signicantly from linear analogues, the branched copolymers exhibited earlier shear-thinning onset in the melt compared 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) 507539

511

synthesized by incorporating the trifunctional comonomer, 1,3,5-tricarboxymethylbenzene [26]. High molecular weight branched aliphatic polyesters such as poly(ethylene succinate), poly(butylene succinate) (PBS), and poly(butylene adipate) (PBA), known as Bionollee polymers, were also prepared [27]. Bionollee polymers are used in a variety of applications, including lm 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., respectively, [30,31]. 2.2. Synthesis of branched polyesters via A2 and B2 monomers 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 the polymerization. Manaresi et al. rst reported the synthesis of branched PETs from dimethyl terephthalate (DMT) and ethylene glycol (EG) using a trifunctional branching agent, trimethyl trimesate (Fig. 2) [19]. To prevent gelation, only small amounts (!2%) of the trifunctional branching agent were used and the inuence 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
O O O O O O O O HO OH O O

COOCH2CH2OH

COOCH2CH2OH

COOCH2CH2OH

HOCH2CH2OOC

COOCH2CH2OH

Fig. 3. Hydroxy ethyl esters formed after the ester-interchange step during the polycondensation of PET.

(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).

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 transesterication step. Transesterication was catalyzed with the addition of manganese acetate at a maximum temperature of 230 8C. Following transesterication, 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 concentrations 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 signicantly inuenced 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

512
O C

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


O O O O O C y
O C

C OCH2CH2O C
x

C OCH2CH2O C

OCH2CH2O C O

C O z

Fig. 4. Structure of randomly branched PET.

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 a Ti(OBu)4 catalyst (Fig. 6) [25]. The branched PBIs were prepared via a two step polycondensation reaction. In the rst 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 bifunctional monomers and trimethyl trimesate (TMT) as a trifunctional comonomer [26]. Using titanium tetrabutoxide 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 inuence 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 adipic acid 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 distribution 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 concentration 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.
O OH

OH OH

HO HO

OH OH

HO

OH O OH O (d)

HO

OH

OH

(a)

(b)

(c)

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) 507539


OH O N C C O N C N O OH

513

O O

O O HO OH HO

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 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 monofunctional 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-benzoylbenzoate, 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 percent conversions, and no gelation occurred when the ratio of monofunctional monomer to trifunctional monomer 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 monofunctional comonomer, and polycondensation temperatures 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 polycondensation reaction in the presence of monofunctional
O HO O O O (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.

(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).

When branched PBAs were prepared with either glycerol or pentaerythritol, it was found that the molecular weight of branched PBAs with pentaerythritol was higher due to the higher degree of functionality. The introduction of a branching agent, TMP, to the polycondensation system of succinic acid and BD resulted in high molecular weight randomly branched poly(butylene succinate) (PBS) [34]. The esterication 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 weights and 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 molecular weight decreased. It is possible to synthesize numerous types of branched polyesters by introducing A3/B3 or A4/B4 monomers into a polymerization of A2 and B2 monomers that involve transesterication and polycondensation. Long chain branched polymers generally have higher weight average molecular weights and broader molecular weight distributions compared with linear polymers synthesized at equivalent reaction 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 inuences thermal, mechanical, and rheological behaviors of polymers as discussed in subsequent sections.

O HO

514

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

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 pentaerythritol 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 polyester molecular weight, while avoiding thermal degradation [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 conducted 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 polymerization step was controlled and specic compositions of reagents were used [23]. The branched PETs were characterized by solution viscometry, thermal analysis, 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 polycondensation 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-tetracarboxylic acid, dipentaerythritol, and tripentaerythritol as
O HO HO O (a) O (b) O OH OH HO OH OH O OH

branching agents with the endcapping reagent benzyl alcohol (Fig. 8). A wide range of branched PETs (with various branching agents) as well as their compositions with and without the presence of an endcapping 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 inuence of branching on solution and melt properties. Although end-group modication 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 Ramakrishnan et al. reported the synthesis and characterization 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 the inuence of kinks, linear and branched polymers
OH OH HO OH OH O OH OH O OH OH OH

(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) 507539

515

were also prepared. Branched and kinked PETs were synthesized using melt polymerization of bis-(2hydroxyethyl) 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) benzoate (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, which resulted 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 as an 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.

516

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

3. Characterization of branched polymers 3.1. Introduction Since branching has such a dramatic inuence 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 difcult 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 inuences polymer molecular weight and molecular weight distribution, it is important to deconvolute the effects of chain architecture 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 (temperature 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, hR2 g i, is a measure of a polymers hydrodynamic volume as measured using static light scattering. Consequently, the ratio of the h R2 g i 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). gZ h R2 g ibranched hR2 g ilinear (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). g0 Z hbranched hlinear (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).
a hlinear Z KMw

(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 of g, denoted gM in the gure, 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 Mark Houwink 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) 507539

517

exponent from 0.73 to 0.68 to 0.39 for linear, starbranched, 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 starshaped 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 branching units for randomly branched chains [48]. For star polymers with monodisperse arms under theta conditions, g can be calculated using Eq. (6), gZ 3f K 2 f2 (6)

where f is 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 1=2 limitation of MALLS. In fact, hR2 measurements gi are unreliable for values less than 10 nm, which is roughly a value of Mw on the order of 104 g/mol [49]. Intrinsic viscosity and g 0 are more reliable measurements, however a theoretical basis is not developed that relates g 0 to molecular parameters since the dependence of g 0 and g is not understood. Much work has focused on understanding this relationship, and empirical correlations suggest the form g 0 Z g3 (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 viscometric 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 methacrylate) with 3270 arms [53]. The authors showed the number of arms, f, was more accurately predicted through the empirical ts 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 corresponding to respective a values of 0.80 and 0.43 mark the average molecular weight between branch points.

518

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

theoretical concepts of Flory and Stockmayer. The models are dependent upon the number of polymer chain ends, molecular weights, and initial concentration of the mono-, bi-, and tri-functional repeat units. 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 endgroups [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, Bn Z 2r 3 K r K 3p (8)

redened as,       M Mt Mb M K r C Mm K b r 0 p Z 1 KE b C 2 3 2 2 (12) where Mm is the molecular weight of the monofunctional unit. Finally, the average number of branches, Bn, is given by Eq. (13). Bn Z 2r 3 K r C 3r 0 K 3p (13)

where r is a parameter that represents the initial polymer composition, rZ 3ntrif 3ntrif C 2nbif (9)

and ntrif is 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),    M Mb Mt K p Z 1 KE b Kr (10) 2 2 3 where E is the sum of all end-groups (equiv/g) and Mb and Mt are the molecular weights of the bi- and trisubstituted repeat units, respectively. A mono-functional agent is often added to a stepgrowth reaction mixture in tandem with the multifunctional 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, r0 Z nmono nmono C 2nbif C 3ntrif (11)

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 aforementioned 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 polyolens, the above methods cannot discriminate between the two thereby making LCB detection difcult. More recently 13C NMR measurements detected branching levels of 0.35 branches per 10,000 carbons in polyethylenes [58]. Since the ow behavior of polymers is sensitive to long chain branches 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. Inuence 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

where nmono is the number of moles of monofunctional agent. The extent of reaction, p, is

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

519

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 branch along a polymer backbone on average is known to signicantly alter the ow properties [60]. Only longchain 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 inuence rheological properties including branch length, degree of branching, and chain architecture (random, starbranched, comb, H-branched, etc.). Typically, branches are introduced in a random fashion during polymerization, thereby leading to a broad distribution of branch lengths and branch density. Consequently it becomes difcult 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 are often 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 applications [62]. This section describes the inuence of branching on the melt and concentrated solution rheological properties of polyesters, with particular focus on the inuence 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 weight. For example, a 100,000 g/mol polymer with an average of one branch per chain has an average branch length of

33,300 g/mol from the relationship, Mb Z Mw 2 Bn C 1 (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 inuences of both parameters on rheological properties. The dependence of h0 on Mw is well established for linear, exible chains. Two regimes are separated by a critical molecular weight (Mc), below which h0 scales directly with Mw and above which h0 generally scales 3:4 with Mw . Chains with molecular weights below Mc are too small to entangle, while the high molecular weight chains are topologically constrained due to entanglement couplings. Researchers have shown a signicant departure from the h0KMw relationship exists for branched chains due to the reduced hydrodynamic volume of the branched chains at low molecular weights, and increased entanglement couplings 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 M w (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 0.1 to 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

520

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

Fig. 12. Systematic decrease in (h0,b/h0,l) as a function of average number of branches per chain for randomly branched PET with MwZ50,000 g/mol.

of h0 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 terephthalate) [54,68]. The inuence of the level of random branching on the rheological properties of aliphatic polyesters has also 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 signicant 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. Lehermeier and Dorgan studied the inuence 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 degradation 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 polylactides were not characterized. Thus, it was difcult 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 without hydrodynamic or topological interactions [72]. Moreover, 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 by analyzing the intrinsic viscosity dependence of Mw as shown in Fig. 11. The Mb (66,000 g/mol) was dened as the crossover from linear to branched behavior, and was measured by the reduction of the 3 :6 [h] vs. Mw slope. Below Mb, h0 scaled with Mw which was consistent with experimental results for 6 :0 entangled linear chains, and above Mb, h0 w Mw 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

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

521

increased from 33,000 for the linear chain to 98,000 and 110,000 g/mol for the trifunctional and tetrafunctional polymers, respectively. Unlike the randomly branched polyesters, the researchers observed a weaker 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, signicantly 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.

the number of entanglements per branch, due to entangled dynamics as the viscosity exponent, s, varies signicantly from the Rouse prediction of 1.33. Consequently, the authors concluded that Mbw2Me for entanglements to dominate the ow 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 rheological 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 Mw dependence of h0 [76]. Moreover, the number of branches per chain also inuenced Mc, as Mc

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.

522

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

observed for Mb/Me!2, due to relaxation of the short backbone segments via Rouse-like motions, while branched chains with M b /M e Z 6 showed h 0 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 welldened 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 commercially 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 specic 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 effect arises 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 conning 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 and relaxation 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 inuenced 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 polymers 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 signicantly 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, respectively. Fig. 15 shows stars with fZ64 and fZ128 arms display an additional transition in the terminal ow range that is not present for stars with fZ4 arms. This additional transition for stars with high degrees of functionality is attributed to cooperative rearrangement 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 and Gruver 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 asymmetric 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) 507539

523

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

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.

length ranged from Mb/MeZ0 to Mb/MeZ18, and was shorter than the backbone length (Mbb/MeZ38). Deviation from linear chain behavior and h0 enhancement were observed for Mb/Mew23, with considerably 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, GZ h0;b M h0:l M (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 endcapped 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 difcult 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.

524

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

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 work has 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 h0 for 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 an increased relaxation time of the cross-bar. Moreover, they showed the terminal relaxation time and h0 enhancement were primarily controlled by the branch

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.

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 (fZ32) [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 inuence of branching on rheological properties, most commercially produced polymers are randomly branched. Interest in the rheological characterization of combbranched 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 backbone molecular 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 compared to linear chains of the same Mw and showed h0 enhancement when compared at equivalent intrinsic viscosity. Surprisingly, this enhancement was not restricted to branch lengths above Me, and the h0 enhancement was different for the combs with different molecular weight backbones. However, the authors found good agreement with the log GKMb/Me relationship 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 polybutadiene 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

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

525

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 Hshaped polymers, marked by relaxation of the arms at intermediate 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 combbranched polybutadiene. Roovers and Toporowski attributed the broader low frequency relaxation to additional couplings between a branch and a backbone 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 polystyrene 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 100 repeat units, a clear plateau region was observed in the dynamic shear modulus. However, the authors did not address the incompatibility issues between polystyrene 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 branched PMMA 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),

 h0 T Z A0 exp

Ea RT

 (16)

where h0 is the zero shear rate viscosity, A0 is a preexponential factor, Ea is the activation energy, R is the gas constant, and T is 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 signicantly 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 lowdensity polyethylene compared to high-density polyethylene. However, limited work has focused on the inuence of branching on Ea in polyesters. Munari et al. investigated the inuence of longchain branching on the ow activation energy for a series 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) compared 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 polyesters and different temperature coefcients of the repeat units. Graessley related this inconsistent behavior to differences in the temperature coefcients 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 retraction or short time-scale uctuations along the tube contour length of an arm. As an arm relaxes, it must pass through a higher energy barrier due to the more

526

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

compact conformational states that are dependent on the temperature coefcient 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 entanglements per branch (Mb/Me), and decrease to zero with decreasing branch length [108]. 5. The inuence 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 polymers 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 considered. 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 interactions are screened at length scales longer than the correlation length, where the correlation length is dened 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 sufcient to topologically constrain the polymer chain motion. Above Ce, the semidilute entangled regime, chain crowding and interpenetration is sufcient 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 dened 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 experimentally measured by analyzing the concentration dependence of specic viscosity (hsp), hsp Z h0 K hs 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 weaker dependence 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 since larger 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 specic 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) 507539

527

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 in poor 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 of hsp and 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 the semidilute 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 exponent decreased from 12.5 to 4.7, attributed to a larger number 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 isophthalate) (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 inuenced 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 bers has received signicant commercial attention, however, electrospinning has recently emerged as a specialized processing technique for the formation of sub-micron, high surface area bers [121]. The utility of branched polyesters in the electrospinning process has received only limited attention. Typically, conventional polymer bers are melt spun using pressure-driven ow through an extruder, yielding bers on the order of 10100 mm in diameter. Electrospinning utilizes an electrostatic potential to create sub-micron size bers on the order of 100 nm in diameter [122,123]. Electrospinning occurs when a charged polymer solution or melt, possessing sufcient molecular entanglements, emits a charged uid jet in the presence of an electric eld. Specically, a charged droplet of partially conductive solution suspended at the end of a capillary deforms into a conical shape, or Taylor cone, when subjected to a Coulumbic force [124]. The Taylor cone is formed due to a balancing of

Fig. 19. Illustration of a branched chain in the semi-dilute regime (COC*).

528

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

the repulsive nature of the charge distribution on the droplets surface and the surface tension of the liquid [125]. As the charge is increased above a critical voltage, a stable jet is discharged from the tip of the Taylor cone. The charged jet follows a chaotic trajectory of stretching and splaying until it reaches a grounded target, thereby completing the circuit. Often, electrospun bers have undesired by-products in the form of beads, which have been extensively studied [126]. A jet of low molecular uid breaks up into small droplets, a phenomenon termed electrospraying, while a polymer solution with sufcient chain overlap and entanglements does not break up, but undergoes a bending instability that causes a whiplike motion between the capillary tip and the grounded target [127]. Researchers have investigated the instabilities of the electrically forced jet and developed a model that relates the dependence of the bending instability on the electric eld and solution owrate [128]. This bending instability accounts for the high degree of single ber drawing that results in submicron size bers. The small ber size and disordered deposition onto the grounded target yield non-woven, three-dimensional ber mats with a high specic surface area and a sub-micron degree of porosity [129]. The unique bers have been suggested for applications such as ltration devices, membranes, vascular grafts, protective clothing, and tissue scaffolds [130132]. The effects of processing variables (applied voltage, tip to target distance, feed rate of the solution to the capillary tip) and solution properties (solvent, viscosity, concentration, conductivity, surface tension) on electrospun ber morphology have been extensively investigated for a variety of polymeric systems [133,134]. For many polymer/ solvent systems, increasing the solution concentration or viscosity decreases the number of bead defects and increases the overall ber diameter of the electrospun bers [135]. Consequently, it was qualitatively shown that chain entanglement is necessary to form defect-free, sub-micron size bers. McKee et al. quantied the importance of chain entanglement on the electrospinning process by spinning solutions of linear and branched PET-co-PEI from the semidilute unentangled and semidilute entangled regime [119]. Fibers formed

from a branched polymer exhibit advantages over bers formed from linear analogs. Branching allows (1) control of chain end concentration for tailored functionalization, (2) controlled degradation for specic drug delivery proles (see Section 6.4), and (3) reduced viscosity for potential melt processing of nanobers. Consequently, it is important to understand the inuence of branching on electrospun ber formation and the rheological/ electrospinning relationships of linear and branched chains. Fig. 20 shows eld emission scanning electron microscopy (FESEM) images of a linear PET-co-PEI (Mw of 77,300 g/mol and CeZ4.5 wt%). Below the Ce of 4.5 wt%, polymer droplets were primarily formed (Fig. 20a). As the concentration was increased above Ce, beaded nanobers were formed with average ber diameters of 0.50 mm (Fig. 20b). When the linear copolyester was electrospun from solutions above 2 times Ce (10 wt%), uniform, defect-free bers were formed with average ber diameters of 1.5 mm (Fig. 20c). As the concentration was increased further to 12 wt% and 17 wt%, uniform bers with nominal diameters of 2.5 and 5 mm were, respectively, electrospun (Fig. 20d and e). Attempts to electrospin 20 wt% solutions of the linear PET-co-PEI were unsuccessful unless the syringe pump owrate was increased to 20 mL/h to overcome the high viscosity of the solution and initiate jet formation. This increase in pump ow rate resulted in the formation of large bers greater than 10 mm in diameter (Fig. 20f). Fiber morphology showed a similar dependence on Ce when electrospun from PET-co-PEI solutions with a range of Mw and g 0 . The concentration dependence of the ber diameter for linear and randomly branched PET-co-PEI were superimposed when the solution concentration was normalized with respect to Ce. Fig. 21 shows the superimposed curve for the dependence of normalized solution concentration (C/Ce) on ber diameter for the PET-co-PEI series, where the error bars represent the ber size distribution. The inuence of chain length and branching architecture on the electrospinning process was removed when the solution concentration was normalized with Ce. The ber diameter scaled with the normalized concentration as,

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

529

Fig. 20. FESEM images of electrospun bers of linear PET-co-PEI (MwZ77,300 g/mol, CeZ4.5 wt%) at several concentrations (3.520 wt%).

 Diam w

C Ce

 2 :6 (18)

6. Inuence of branching on thermal properties of polyesters 6.1. Introduction The incorporation of branching into polymers dramatically inuences the thermal properties, such as the melting temperature (T m) if molecular symmetry allows crystallization, glass transition

where C is the polymer concentration in solution and Ce is the entanglement concentration. It should be noted that Eq. (18) is only valid for electrospinning PET-co-PEI from a 70/30 CHCl3/DMF w/w cosolvent at room temperature since Ce is dependent on hydrodynamic dimensions.

530
10

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

1
Diam[m] = 0.18 R 2 = 0.95

( e)
C C

2.6

0.1 1

C/Ce

10

Fig. 21. Dependence of ber diameter on normalized concentration for linear and branched PET-co-PEI.

domain of the experiment. Below the Tg, the polymer lacks mobility, but maintains the disordered state of the melt. As an amorphous polymer is cooled from the melt its free volume decreases since the thermal energy for chain mobility decreases. The Tg occurs, once the free volume shrinks such that cooperative motion of the backbone is prohibited. Thus, any variables that inuence the polymers fractional free volume can affect the Tg. Some examples include, chemical composition, molecular weight, crystallinity, addition of plasticizer, and chain topology. The glass transition temperature for random copolymers often can be reasonably well described with chemical composition by the Fox Equation, 1 w 1 K wA Z A C Tg Tg;A Tg;B (19)

temperature (Tg), degree and rate of crystallization, and the isotropic transition temperature (Ti) for liquid crystalline polymers Short-chain branches affect solid-state properties and limit chain crystallization, while long chain branches generally enhance h0 and provide shear thinning with longer relaxation times. For high performance applications, long chain branching is utilized to control the rheological and processing properties, while short chain branching inuences thermal behavior and mechanical properties. Often, short chain branching is introduced to reduce the melting transition and consequently improve the melt processibilty of a polymer. It should be noted that for step-growth polymers it is often difcult to separate effects from comonomer incorporation and branching on thermal properties. For example, generally branched polyester synthesis utilizes a comonomer with functionality greater than two that is added to the polymerization to yield a branched structure. Moreover, thermal transitions and crystallization behavior are dependent on both comonomer composition and degree of SCB or LCB. Consequently, due to the nature of random branching in polyesters chemical composition and degree of SCB or LCB must be accounted for. 6.2. Glass transition The glass transition temperature (Tg) is the temperature at which a polymer chain possesses sufcient thermal energy that cooperative, segmental motion of the backbone occurs within the frequency

Avg. Diam. (m)

where wA and wB are the respective weight fractions of monomers A and B in the copolymer, and Tg,A and Tg,B are the homopolymer glass transition temperatures. Thus, the Tg of a randomly branched polyester is a function of the multi-functional (fO2) branching agent composition, and must be considered when addressing the inuence of branching on Tg. As a specic example of this issue, Jayakannan and Ramakrishnan attempted to separate the effects of copolymer composition and branching by synthesizing a series of linear, kinked, and branched PET samples [24]. Fig. 9 shows the structures of the three PETs, where the kinked PET has a linear architecture but contains the same 1,3 connectivity as the branched PET. Fig. 22 shows the Tg of PET was unaffected by either the 1,3 connectivity or the branches at a range of compositions. In addition to PET, the independence of Tg with branching was observed for poly(butylene isophthalate) [25]. Another important issue with investigating the inuence of chain architecture on the glass transition temperature is separating the effects of short chain branches and long chains branches. In the current literature there is a lack of structure/property relationships for well-dened polyesters with known compositions of short chain and long chain branches. However, thermal studies performed on short chain branched poly(n-alkyl methacrylates) with varying alkyl chain length can be applied to polyesters. As the length of the alkyl side chain is increased from

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

531

Fig. 22. DSC trace for branched PET with 0 to 5 mol% branching comonomer.

1 carbon in poly(methyl methacrylate) to 12 carbons in poly(dodecyl methacrylate), the Tg systematically decreased by ca. 190 8C [136]. Reductions in Tg have been clearly correlated to increasing fractional free volume with alkyl length. Unlike, the poly(n-alkyl methacrylates) with a side chain on every repeat unit of the backbone, a series of short chain branched aliphatic polyesters with varying compositions of SCB were studied [137]. The Tg of poly(butylene succinate) (PBS) with ethyl, n-hexenyl, and octyl branches systematically decreased with the degree of SCB. Yoon et al. also showed a similar decrease in Tg with short chain branching content in poly(ethylene adipate) and other aliphatic polyesters [138]. Up to this point, only the role of SCB on Tg has been discussed. If molecular symmetry exists and the side chains are long enough to order, side-chain crystallization may occur which increases the Tg [139]. This phenomenon is due to organization and packing of the side chains into a rigid crystalline structure. Greenberg and Alfrey observed side chain crystallization for a series of poly(n-alkyl methacrylates) and poly(n-alkyl acrylates) with n-alkyl groups from 12 to 18 carbons long [140]. The side chain crystalline domains serve to restrict chain segment mobility in the amorphous region, which subsequently broadens and raises the glass transition temperature [141]. The crystalline regions act as physical crosslinks where chains from the amorphous region are

physically tied to the crystals. Further discussion of side chain crystallization is addressed in Section 6.3. Although the role of comonomer content and short chain branches on Tg has been studied for poly(n-alkyl methacrylates) and alkyl substituted rigid polymers, little work has focused on how these variables inuence the thermal properties of semi-exible polyesters [142,143] Careful studies on well-dened branched polyesters is needed to ll this gap in the literature. 6.3. Melting behavior and quiescent crystallization growth Polymer chains that contain molecular symmetry typically crystallize under appropriate conditions. The crystalline content of polymers is always less than 100% and the crystalline regions exist either as a matrix containing the amorphous domains or within in an amorphous matrix [144]. The individual crystals usually are preferentially comprised of folded chain lamellae, which are approximately 10 nm thick with polymer chains folded back on themselves. Amorphous tie chains interconnect the lamellae with each other, and lamellae structures that align radially form a spherulitic morphology for systems crystallized from the oriented state [145]. In quiescent crystallization, crystal growth proceeds from a nucleation process. Nucleation can occur

532

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

due to random uctuations of macromolecules that result in favorable alignment of the polymer chains called homogeneous nucleation. Nucleation more commonly occurs at the interface of a second phase, which is termed heterogeneous nucleation. For a symmetric polymer that can crystallize, crystallization growth occurs at a temperature between Tg and Tm. Above Tm crystalline regions melt, and below Tg the chain does not possess sufcient segmental mobility to undergo nucleation. Consequently, TmKTg denes the window for crystallization for chains with a single melting transition, and polymers with broader windows generally crystallize more readily. For example, linear polyethylene that has a highly symmetric backbone and crystallizes very rapidly has a TmKTg of approximately 140 8C, while bisphenol-A polycarbonate, a polymer very slow to crystallize, has a window of only ca. 100 8C. Long-chain branching does not have a large inuence on the Tg, while short chain branches often depress the Tg due to increases in free volume as discussed in Section 6.2. However, both SCB and LCB have a signicant effect on the level of crystallinity and crystalline melting. For example, side chain branching in poly(a-olen) homopolymers inuences Tm through entropic considerations. The Tm increases as the side chain increases from 0 carbons (polyethylene) to 1 carbon (polypropylene) due to steric crowding of the backbone from the methyl group which inuences the chain conformation in the melt [144]. As the linear side chain is increased further from 1 carbon to 4 carbons long, the melting transition decreases. In this case the longer side chains have increased exibility thereby allowing a larger number of chain conformations in the melt, which results in a depression of Tm. The Tm reaches a minimum at a branch length of 4 carbons and then increases again due to ordering and crystallization of the side chains [146]. In fact, as the branch length is greatly extended to molecular weights above Me (LCBs), crystallization of the branch usually occurs. Up to this point the discussion has focused on isotactic short-chain branched homopolymers, which contained a branch on every repeat unit. However, often SCBs are introduced into a polymer backbone in a random fashion by copolymerization. The most widely studied and classic example is linear lowdensity polyethylene (LLDPE). Copolymerizing

ethylene with 1-butene, 1-hexene, or 1-octene results in LLDPE with branches of 2, 4, or 6 carbons, respectively. Alamo et al. reported a decrease in Tm with SCB content for a series of ethylene copolymers due to disruption of local chain symmetry [147]. However, the authors observed Tm was independent of the comonomer type, which suggested the three branch lengths were equally effective at disrupting crystallinity and did not become incorporated into the lattice. It is important to note that for the case of branching in polyesters it becomes more difcult to separate effects of branching and comonomer incorporation on Tm and crystallization growth since a multi-functional comonomer (fO2) is used to induce branching. Moreover, in typical step-growth polymerizations it is very difcult to determine exact branch length or even branch type (SCB vs. LCB). The role of comonomer incorporation, chain extension, and crosslinking of linear PETs to tailor the crystallization kinetics has been extensively studied [148150]. Several researchers have also explored the role of branching on thermal transitions and crystallization kinetics of PET [151153] and other polyesters [25,138] In general, a depression and broadening of Tm coupled with lower degrees of crystallinity was observed with increasing branching content due to disruption of chain symmetry. However, the polyesters were not well dened and there was no characterization of branch length or branch type. Consequently, careful thermal and crystallization studies performed on branched polyesters with controlled branch placement and branch length would be a valuable contribution to this area. 6.4. Controlling the biodegradation of aliphatic polyesters through branching Biodegradable polyesters are utilized in environmental plastics and biomedical applications. In particular, aliphatic polyesters based on lactic acid and/or glycolic acid are used for controlled drug delivery release, surgical sutures, prosthetics, and bone screws [154]. The biodegradability of a polymer chain is dependent on the backbone structure and chemical composition [28]. The rate of aliphatic polyester degradation is largely inuenced by the hydrophilicity of the repeat unit, molecular weight, crystallinity, and the polymer environment [155]. As

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

533

Fig. 23. Schematic of polylactide combs with varying backbone molecular weight, arm molecular weight and number of arms.

discussed in the Section 6.3, short chain and long chain branching greatly affect the crystalline content of aromatic and aliphatic polyesters. Much research has focused on tailored biodegradation by means of manipulating polyester chain architecture. The introduction of low amounts of comb shaped branches and hydrophilic units signicantly decreased the crystallinity of poly(L-lactide) (PLLA) [156]. Kim et al. studied the biodegradation of unoriented n-octyl and ethyl branched PEA and PBS [157]. It was observed that degradation rates were fastest for n-octyl branched PEA, while rates were relatively slower for both ethyl and n-octyl branched PBS due to reduced crystallinity of the former. Tasaka et al. synthesized various comb-shaped poly(L-lactide)s with different architectures for controlled degradation for potential biomedical applications [158]. Fig. 23 shows a schematic of the PLLA combs with varying backbone molecular weight, arm molecular weight, and number of arms. The researchers observed the degradation rate was not only dependent on the percent crystallinity, but also the molecular architecture. Among the three different combs shown in Fig. 23 the degradation rate increased signicantly with the number of arms. It should be noted that crystalline content decreased from 22 to 14.6% (as measured by DSC) with increasing number of arms, which would inuence the PLLA degradation as well. For the controlled delivery of bioactive agents, it is often ideal to have uniform, continuous release patterns [159]. Typically, drug release of peptides and proteins from linear polyesters yields discontinuous release proles [160]. Generally, proteins are

impregnated into polyester microspheres through a emulsion technique [161]. Researchers have modied aliphatic polyesters by introducing hydrophilic groups and branch points to circumvent the discontinuous release proles [162]. Breitenbach, Li, and Kissel reported the dependence of degradation kinetics and drug release patterns on the branching architecture from a series of star copolymers with polyethylene oxide (PEO) inner block and either a PLLA or poly(lactic-co-glycolic acid) (PLG) outer block [163]. Both of the block star copolymers showed a more uniform release of an antibiotic during its degradation compared to linear PEO-PLLA and PEO-PLG block copolymers. The improved release prole was attributed to the reduced crystallinity of the star block copolymers. Fig. 24 compares the

Fig. 24. Comparison of the release prole of dextran from linear and star branched poly(glycolic acid-b-ethylene oxide) microspheres.

534

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

release of dextran from microspheres prepared from linear and 4-arm star poly(glycolic acid-b-ethylene oxide). This uniform antibiotic release was attributed to slower initial degradation of the PEO blocks due to the larger number of connecting bonds between PEO and PLG. Moreover, comb-branched polyesters with PLG arms grafted onto a hydrophilic, polyelectrolyte backbone also showed enhanced delivery proles. 7. Conclusions and future directions This review outlined the inuence of molecular architecture on the properties of polyesters, with the emphasis on both the effect of short and long chain branching on the melt rheological, thermal, and crystallization behavior. A detailed summary of step-growth polymerization methodologies to synthesize randomly branched polyesters was provided. In particular, controlling the ratio of mono-functional to multi-functional (fO2) monomers provides the ability to form high molecular weight, long-chain branched polyesters without gelation. The advantages and disadvantages of various analytical methods to determine the degree of branching and the branch molecular weight, Mb, were reviewed, including dilute solution methods, end-group analysis, and spectroscopic techniques. The majority of the review focused on the melt and solution rheological behavior and included some features of the solid-state properties of both branched and linear polyesters. Shortchain branches affect solid-state properties and limit chain crystallization, while long chain branches (MbOMe) generally enhance h0 and provide early shear thinning onset due to longer relaxation times in the melt and concentrated solutions. For high performance applications, long chain branching is utilized to control the rheological and processing properties, while short chain branching inuences thermal behavior and mechanical properties. While the majority of this outline focused on the synthesis, rheological and solid-state properties of branched polyesters, some very limited discussion was devoted to the inuence of branching on ber formation and controlled biodegradation for drug-delivery applications. As mentioned previously, little work has been performed on how SCB or LCB in polyesters affects thermal behavior. Thus, careful studies need to be performed on well-dened branched polyesters,

where the branch length and branching composition are known exactly.

Acknowledgements This material is based upon work supported by the US Army Research Laboratory and US Army Research Ofce under grant number DAAD19-02-10275 Macromolecular Architecture for Performance (MAP) MURI. The authors also thank Eastman Chemical Company for their continued support of polyester research.

References
[1] Roovers J. In: Kroschwithz JI, editor. Encyclopedia of polymer science and technology. New York: Wiley; 1985. p. 4789. [2] Quirk RP, Lee Y, Kim J. In: Mishra MK, Kobayashi S, editors. Star and hyperbranched polymers. New York: Marcell Dekker; 1999. p. 125. [3] Shroff RN, Mavridis H. Long-chain branching index for essentially linear polyethylenes. Macromolecules 1999;32: 845464. [4] Pitsikalis M, Pispas S, Mays JW, Hadjichristidis N. Effect of block copolymer architecture. Adv Polym Sci 1998;135:1. [5] Tande BM, Wagner NJ, Mackay ME, Hawker CJ, Jenng M. Viscometric, hydrodynamic, and conformational properties of dendrimers and dendrons. Macromolecules 2001;34: 85803. [6] Hult A, Malmstrom E, Johansson M. In: Salamone JC, editor. Polymeric materials encylopedia. New York: CRC Press; 1996. p. 31717. [7] Lin Q, Long TE. Polymerization of A2 with B3 monomers: a facile approach to hyperbranched poly(aryl ester)s. Macromolecules 2003;36:980816. [8] Malmberg A, Kokko E, Lofgren B, Seppala JV. Long-chain branched polyethylene polymerized by metallocene catalysts Et[Ind]2ZrCl2/MAO and Et[IndH4]2ZrCl2/MAO. Macromolecules 1998;31:844854. [9] Hsieh T-T, Tiu C, Hsieh K-H, Simon GP. Characterization of thermotropic liquid crystalline polyester/polycarbonate blends: miscibility, rheology, and free volume behavior. J Appl Polym Sci 2000;77:231930. [10] Bikiaris DN, Karayannidis GP. Synthesis and characterization of branched and partially crosslinked poly(ethylene terephthalate). Polym Int 2003;52:12309. [11] Han Y-K, Um JW, Im SS, Kim BC. Synthesis and characterization of high molecular weight branched PBA. J Polym Sci Part A: Polym Chem 2001;39:214350. [12] Anderson KS, Hillmyer MA. Melt chain dimensions of polylactide. Macromolecules 2004;37:185762.

M.G. McKee et al. / Prog. Polym. Sci. 30 (2005) 507539 [13] Ihn KJ, Yoo ES, Im SS. Structure and morphology of poly(tetramethylene succinate) crystals. Macromolecules 1995;28:24604. [14] Ferry JD. Viscoelastic properties of polymers. 3rd ed. New York: Wiley; 1980. [15] Fetters LJ, Lohse DJ, Graessley WW. Chain dimensions and entanglement spacings in dense macromolecular systems. J Polym Sci Part B: Polym Phys 1999;37:102333. [16] Zang Y-H, Carreau PJ. A correlation between critical end-toend distance for entanglements and molecular chain diameter of polymers. J Appl Polym Sci 1991;42:19658. [17] Hudson H, MacDonald WA, Neilson A, Richards RW, Sherrington DC. Synthesis and characterization of nonlinear PETs produced via a balance of branching and Endcapping. Macromolecules 2000;33:925561. [18] Yoon KH, Min BG, Park OO. Effect of multifunctional comonomers on the properties of poly(ethylene terephthalate) copolymers. Polym Int 2002;51:1349. [19] Manaresi P, Parrini P, Semeghini GL, de Fornasari E. Branched poly(ethylene terephthalate) correlations between viscosimetric properties and polymerization parameters. Polymer 1976;17:595600. [20] Nagata M. Synthesis, characterization, and enzymatic degradation of novel regular network aliphatic polyesters based on pentaerythritol. Macromolecules 1997;30: 652530. [21] Neff BL, Overton JR. Characterization of branched copolyesters using gel permeation chromatography with on-line low anlge light scattering. ACS Polym Prepr 1982;23:130. [22] Manaresi P, Munari A, Pilati F, Alfonso GC, Russo S, Sartirana L. Synthesis and characterization of highly branched poly(ethylene terephthalate). Polymer 1986;27: 95560. [23] Rosu RF, Shanks RA, Bhattacharya SN. Shear rheology and thermal properties of linear and branched poly(ethylene terephthalate) blends. Polymer 1999;40:58918. [24] Jayakannan M, Ramakrishnan S. Synthesis and thermal analysis of branched and kinked poly(ethylene terephthalate). J Polym Sci Part A: Polym Chem 1998;36: 30917. [25] Finelli L, Lotti N, Munari A. Inuence of branching on the thermal behavior of poly(butylene isophthalate). J Appl Polym Sci 2002;84:200110. [26] Righetti MC, Munari A. Inuence of branching on melting behavior and isothermal crystallization of poly(butylene terephthalate). Macromol Chem Phys 1997;198:36378. [27] Fujimaki T. Processability and properties of aliphatic polyesters BIONOLLE, synthesized by polycondensation reaction. Polym Degr Stab 1998;59:20914. [28] Yoshikawa K, Ofuji N, Imaizumi M, Moteki Y, Fujimaki T. Molecular weight distribution and branched structure of biodegradable aliphatic polyesters determined by s.e.c.MALLS. Polymer 1996;7:12814. [29] Kumar A, Ramakrishnan S. Effect of branching and molecular kinks on the properties of main chain thermotropic liquid crystalline polymers containing exible spacers. Macromolecules 1996;29:85513.

535

[30] Kricheldorf HR, Luubers D. New polymer synthesis. 81. Poly(3-oxybenzoate) randomly branched with 3,5-dihydroxybenzoic acid or 5-hydroxyisophalic acid. Macromol Chem Phys 1995;196:15493562. [31] Jayakannan M, Rammakrishnan S. Effect of branching on the thermal properties of novel branched poly(4-ethyleneoxy benzoate). J Polym Sci Part A: Polym Chem 2000;38:2618. [32] Weisskopf K. Characterization of poly(ethylene terephthalate) by gel permeation chromatography. J Appl Polym Sci 1990;39:214152. [33] Hess C, Hirt P, Opperman W. Inuence of branching on the properties of poly(ethylene terephthalate) bers. J Appl Polym Sci 1999;74:72834. [34] Kim EK, Bae JS, Im SS, Kim BC, Han YK. Preparation and properties of branched polybutylenesuccinate. J Appl Polym Sci 2001;80:138894. [35] Wu D, Chen F, Li R. Reaction kinetics and simulations for solid-state polymerization of poly(ethylene terephthalate). Macromolecules 1997;30:673742. [36] Karayannidis GP, Kokkalas E, Bikiaris DN. Solid-state polycondensation of poly(ethylene terephthalate) recycled from postconsumer soft-drink bottles. II. J Appl Polym Sci 1995;56:40510. [37] Tate S, Ishimaru F. Polymer. Swollen-state polymerization of poly(ethylene terephthalate): kinetic analysis of reaction rate and polymerization conditions 1995;36:3536. [38] Oh SJ, Kim BC. Effects of hydroxyl-group end capping and branching on the physical properties of tailored polyethylene terephthalates. J Polym Sci Part B: Polym Phys 2001;39: 102735. [39] Kratochivil P. Characterization of branched polymers. Macromol Symp 2000;152:27987. [40] Zimm BH, Kilb RW. Dynamics of branched polymer molecules in dilute solution. J Polym Sci Part B: Polym Phys 1996;34:136790. [41] Podzimek S. A review on the analysis of branched polymers by SEC-MALS. American Laboratory January 2002;3845. [42] Grcev S, Schoenmakers P, Iedema P. Determination of molecular weight and size distribution and branching of PVAc by means of size exclusion chromatography/multiangle laser light scattering (SEC/MALLS). Polymer 2003;45: 3948. [43] Kharchenko SB, Kannan RM, Cernohous JJ, Venkataramani S. Role of architecture on the conformation, rheology, and orientation behavior of linear, star, and hyperbranched polymer melts.1. Synthesis and molecular characterization. Macromolecules 2003;36:399406. [44] Kharchenko SB, Kannan RM. Role of architecture on the conformation, rheology, and orientation behavior of linear, star, and hyperbranched polymer melts.2. Linear viscoelasticity and ow birefringence. Macromolecules 2003;36: 40715. [45] Roovers J, Toporowski P, Martin J. Synthesis and characterization of multiarm star polybutadienes. Macromolecules 1989;22:1897903. [46] Bauer BJ, Fetters LJ, Graessley WW, Hadjichristidis N, Quack GF. Chain dimensions in dilute polymer solutions:

536

M.G. McKee et al. / Prog. Polym. Sci. 30 (2005) 507539 a light-scattering and viscometric study of multiarmed polyisoprene stars in good and q solvents. Macromolecules 1989;22:233747. Lusignan CP, Mourey TH, Wilson JC, Colby RH. Viscoelasticity of randomly branched polymer in the vulcanization class. Phys Rev E 1999;60:565769. Zimm BH, Stockmayer WH. The dimensions of chain molecules containing branches and rings. J Chem Phys 1949;17:130114. de Luca E, Richards RW. Molecular characterization of a hyperbranched polyester.1.Dilute solution properties. J Polym Sci Part B: Polym Phys 2003;41:133951. Douglas JF, Roovers J, Freed KF. Characterization of branching architectures through universal ratios of polymer solution properties. Macromolecules 1990;23:416880. Bohdanecky M. A semiempirical formulation of the effect of random branching on intrinsic viscosity. Macromolecules 1977;10:9715. Jackson C, Chen Y-J, Mays JW. Dilute solution properties of randomly branched poly(methyl methacrylate). J Appl Polym Sci 1996;59:17988. Balke ST, Mourey TH, Robello DR, Davis TA, Kraus A, Skonieczny K. Quantitative analysis of star-branched polymers by multidetector size-exclusion chromatography. J Appl Polym Sci 2002;85:55270. Munari A, Pezzin G, Pilati F, Manaresi P. Rheological characterization of highly branched poly(ethylene terephthalate). Rheol Acta 1989;28:259. Ahmad NM, Lovell PA, Underwood SM. Viscoelastic properties of branched polyacrylate melts. Polym Int 2001; 50:62534. Doerpinghaus PJ, Baird DG. Separating the effects of sparse long-chain branching on rheology from those due to molecular weight in polyethylene. J Rheol 2003;47:71736. Colby RH, Janzen J. Diagnosing long-chain branching in polyethylenes. J Mol Struct 1999;485-486:56984. DeGrott AW, Karjala TP, Taha AN, Johnson MS. Longchain branching measurements of polyethylenes: a tool-box of techniques. International workshop on branched polymers for performance, Proceedings. Spring 2004. Wood-Adams P, Dealy JM, deGroot AW, Redwine GD. Effect of molecular structure on the linear viscoelastic behavior of polyethylene. Macromolecules 2000;33: 748999. Bin Wadud SE, Baird DG. Shear and extensional rheology of sparsely branched metallocene-catalyzed polyethylene. J Rheol 2000;44:115167. Graessly WW. Viscoelasticity and ow in polymer melts and concentrated solutions. Physical properties of polymers 1984 p. 97153. Dhavalikar R, Yamaguchi M, Xanthos M. Molecular and structural analysis of a triepoxide-modied poly(ethylene terephthalate) from rheological data. J Polym Sci Part A: Polym Chem 2003;41:95869. Kang H, Lin Q, Armentrout SR, Long TE. Synthesis and characterization of telechelic poly(ethylene terephthalate) sodiosulfate ionomers. Macromolecules 2002;35:873844. [64] Yilmazer U, Xanthos M, Bayram G, Tan V. Viscoelastic characteristics of chain extended/branched and linear polyethylene terephthalate resins. J Appl Polym Sci 2000;75: 13717. [65] Lin Q, Unal S, Fornof AR, Wei Y, Huimin L, Armentrout SR, Long TE. Synthesis and characterization of poly(ethylene glycol) methyl ether end-capped poly(ethylene terephthalate). Macromol Symp 2003;199:16372. [66] Rosu RF, Shanks RA, Bhattacharya SN. Dynamic rheology of branched poly(ethylene terephthalate). Polym Int 2000;49: 2038. [67] Munari A, Pilati F, Pezzin G. Melt viscosity of linear and branched poly(butylene isophthalate). Rheol Acta 1988;27: 1459. [68] Munari A, Pilati F, Pezzin G. Linear and branched poly(butylene terephthalate): activation energy for melt ow. Rheol Acta 1985;24:5346. [69] Yan D, Wang W-J, Zhu S. Effect of long chain branching on rheological parameters of metallocene polyethylene. Polymer 1999;40:173744. [70] Jin H-J, Park J-K, Park K-H, Kim M-N, Yoon J-S. Properties of aliphatic polyesters with n-parafnic side branches. J Appl Polym Sci 2000;77:54755. [71] Lehermeier HJ, Dorgan JR. Melt rheology of poly(lactic acid): Consequences of blending chain architectures. Polym Eng Sci 2001;41:217284. [72] Lusignan CP, Mourey TH, Wilson JC, Colby RH. Viscoelasticity of randomly branched polyesters in the critical percolation class. Phys Rev Part E 1995;52:627180. [73] Colby RH, Gillmor JR, Rubinstein M. Dynamics of nearcritical polymer gels. Phys Rev Part E 1993;48:37126. [74] Long VC, Berry GC, Hobbs LM. Solution and bulk properties of branched polyvinyl acetates IV-Melt viscosity. Polymer 1964;5:51724. [75] Fetter LJ, Lohse DJ, Milner ST, Graessley WW. Packing length inuence in linear polymer melts on the entanglement, critical, and reptation molecular weights. Macromolecules 1999;32:684751. [76] Valles EM, Macosko CW. Structure and viscosity of poly(dimethylsiloxanes) with random branches. Macromolecules 1979;12:5216. [77] Masuda T, Ohta Y, Onogi S. Rheological properties of randomly branched polystyrenes with different molecular weights between branch points. Macromolecules 1986;19: 252432. [78] Hatzikiriakos SG, Kazatchkov IB, Vlassopouolos D. Interfacial phenomena in the capillary extrusion of metallocene polyethylenes. J Rheol 1997;41:1299316. [79] Pfeiffer G, Gorda KRJ. Star-shaped condensation polymers: synthesis, characterization, and blend properties. J Appl Polym Sci 1993;50:197783. [80] McLeish TCB, Milner ST. Entangled dynamics and melt ow of branched polymers. Adv Polym Sci 1999;43: 195256. [81] Ye X, Sridhar T. Shear and extensional properties of threearm star polystyrene solutions. Macromolecules 2001;34: 82707.

[47]

[48]

[49]

[50]

[51]

[52]

[53]

[54]

[55]

[56]

[57] [58]

[59]

[60]

[61]

[62]

[63]

M.G. McKee et al. / Prog. Polym. Sci. 30 (2005) 507539 [82] Birshtein TM, Zhulina EB, Borsov OV. Temperatureconcentration diagram for a solution of star-branched macromolecules. Polymer 1986;27:107886. [83] Kraus G, Gruver JG. Rheological properties of multichain polybutadienes. J Polym Sci Part A: Polym Chem 1965;3: 10522. [84] Fetters LJ, Kiss AD, Pearson DS, Quack GF, Vitus FJ. Rheological behavior of star-shaped polymers. Macromolecules 1993;26:64754. [85] Klein J. Dynamics of entangled linear, branched, and cyclic polymers. Macromolecules 1986;19:10518. [86] Pakula T, Vlassopoulos D, Fytas G, Roovers J. Structure and dynamics of melts of multiarm polymer stars. Macromolecules 1998;31:893140. [87] Pakula T. Static and dynamic properties of computer simulated melts of multiarm polymer stars. Comput Theor Polym Sci 1998;8:2130. [88] Gell CB, Graessley WW, Efstratiadis V, Pitsikalis M, Hadjichristidis N. Viscoelasticity and self-diffusion in melts of entangled asymmetric star polymers. J Polym Sci Part B: Polym Phys 1997;35:194354. [89] Dorgan JR, Williams JS, Lewis DN. Melt rheology of poly(lactic acid): entanglements and chain architecture effects. J Rheol 1999;43:114155. [90] Milner ST, McLeish TCB. Parameter-free theory for stress relaxation in star polymer melts. Macromolecules 1997;30: 215966. [91] Claesson H, Malmstrom E, Johansson M, Hult A. Synthesis and characterization of star branched polyesters with dendritic cores and the effect of structural variations on zero shear rate viscosity. Polymer 2002;43:35118. [92] Claesson H, Malmstrom E, Johansson M, Hult A, Doyle M, Manson J-AE. Rheological behavior during UV-curing of a star-branched polyester. Prog Org Coatings 2002;44:637. [93] Roovers J. Melt rheology of H-shaped polystyrenes. Macromolecules 1984;17:1196200. [94] Archer LA, Varshney SK. Synthesis and relaxation dynamics of multiarm polybutadiene melts 1998;31:634855. [95] Houli S, Iatrou H, Hadjichristidis N, Vlassopoulos D. Synthesis and viscoelastic properties of model dumbbell copolymers consisting of a polystyrene connector and two 32-arm star polybutadiene 2002;35:65927. [96] Simon PF, Muller AH, Pakula T. Characterization of highly branched poly(methyl methacrylate) by solution viscosity and viscoelastic spectroscopy. Macromolecules 2001;34: 167784. [97] Noda I, Horikawa T, Kato T, Fujimoto T, Nagawasa M. Solution properties of comb-shaped polystyrenes. Macromolecules 1970;3:7959. [98] Roovers J, Graessley WW. Melt rheology of some model comb polystyrenes 1981;14:76673. [99] Daniels DR, McLeish TCB, Crosby BJ, Young RN, Fernyhough CM. Molecular rheology of comb polymer melts. 1. Viscoelastic Response 2001;34:702533. [100] McLeish TCB, Allgaier J, Bick DK, Bishko G, Biswas P, Blackwell R, Biottiere B, Clarke N, Gibbs B, Groves DJ, Hakiki A, Heenan RK, Johnson JM, Kant R, Read DJ,

537

[101] [102]

[103]

[104] [105]

[106]

[107]

[108]

[109] [110]

[111]

[112]

[113]

[114]

[115]

[116]

[117] [118]

Young RN. Dynamics of entangled H-polymers: theory, rheology, and neutron scattering. Macromolecules 1999;32: 673458. Roovers J, Toporowski PM. Relaxation by constraint release in combs and star-combs. Macromolecules 1987;20:23006. Namba S, Tsukahara Y, Kaeriyama K, Okamoto K, Takahashi M. Bulk properties of multibranched polystyrenes from polystyrene macromonomers: rheological behavior I. Polymer 2000;41:516571. Tsukahara V, Namba S, Iwasa J, Nakano Y, Kaeriyama K, Takahashi M. Bulk properties of poly(macromonomers) of increased backbone and branch lengths. Macromolecules 2001;34:26249. Severs ET. Rheology of polymers. New York: Reinhold; 1962. Wilkes GL. An overview of the basic rheological behavior of polymer uids with an emphasis on polymer melts. J Chem Ed 1981;58:88092. Munari A, Pezzin G, Pilati F. Linear and branched poly(butylene terephthalate): activation energy for melt ow. Reol Acta 1990;20:46974. Graessley WW. Effect of long branches on the temperature dependence of viscoelastic properties in polymer melts. Macromolecules 1982;15:11647. Raju VR, Rachapudy H, Graessley WW. Properties of amorphous and crystallizable hydrocarbon polymers. J Polym Sci Part B: Polym Phys 1979;17:122335. Boris DC, Colby RH. Rheology of sulfonated polystyrene solutions. Macromolecules 1998;31:574655. Dobrynin AV, Colby RH, Rubinstein M. Scaling theory of polyelectrolyte solutions. Macromolecules 1995;28: 185971. Colby RH, Fetters LJ, Funk WG, Graessley WW. Effects of concentration and thermodynamic interaction on the viscoelastic properties of polymer solutions. Macromolecules 1991; 24:387382. Graessley WW. Polymer chain dimensions and the dependence of viscoelastic properties on concentration, molecular weight, and solvent power. Polymer 1980;21:25862. Krause WE, Bellomo EG, Colby RH. Rheology of sodium hyaluronate under physiological conditions. Biomacromolecules 2001;2:659. Takahashi V, Isono Y, Noda I, Nagawasa M. Zero-shear viscosity of linear polymer solutions over a wide range of concentration. Macromolecules 1985;18:10028. Isono Y, Nagawasa M. Solvent effects on rheological properties of polymer solutions. Macromolecules 1980;13: 8627. Colby RH, Rubinstein M. Two-parameter scaling for polymer in theta solvents. Macromolecules 1990;23: 27537. Burchard W. Solution properties of branched macromolecules. Adv Polym Sci 1999;143:11194. Sendijarevic I, Liberature MW, McHugh AJ. Effect of branching on the rheological properties of solutions of aromatic etherimide copolymers. J Rheol 2001;45: 124558.

538

M.G. McKee et al. / Prog. Polym. Sci. 30 (2005) 507539 [136] Everaerts AI, Clemens LM. Pressure sensitive adhesives. In: Chaudhury M, Pocius AV, editors. Surfaces, chemistry, and applications. Amsterdam: Elseveir; 2002. p. 487. [137] Jin H-J, Kim K-S, Kim M-N, Lee I-M, Lee H-S, Yoon J-S. Synthesis and properties of poly(butylene succinate) with nhexenyl side branches. J Appl Polym Sci 2001;81:221926. [138] Jin H-J, Park J-K, Park K-H, Kim M-L, Yoon J-S. Properties of aliphatic polyesters with n-parafnic side branches. J Appl Polym Sci 2000;77:54755. [139] Lee JL, Pearce EM, Kwel TK. Side-chain crystallization in alkyl substituted semiexible polymer. Macromolecules 1997;30:687783. [140] Greenberg SA, Alfrey T. Side chain crystallization of n-alkyl polymethacrylates and polyacrylates. J Am Chem Soc 1954; (76):62805. [141] Nogales A, Ezquerra TA, Batallan F, Frick B, LopezCabarcos E, Balta-Calleja FJ. Restricted dynamics of poly(ether ether ketone) as revealed by incoherent quasielastic neutron scattering and broad-band dielectric spectroscopy. Macromolecules 1999;(32):23018. [142] Kricheldorf JH, Domschke A. Layer structures. 3. Poly(hydroquinone-terephthalate)s with one, two, or four alkyl substituents: thermotropic and isotropic rigid rods. Macromolecules 1996;29:133744. [143] Kricheldorf JH, Domschke A. Layer structures. 1. Poly(phenylene-terephthalate)s derived from mono-, di-, and tetrakis(alkylthio)terephthalic acids. Macromolecules 1994; 27:150916. [144] Mandelkern L. Crystallization of polymers. Volume 1 Equilibrium concepts. Cambridge: Cambridge University Press; 2002. [145] van Krevelen. Crystallinity of polymers and the means to inuence the crystallization process. Chimia 1978;32: 27994. [146] Trafara G, Koch R, Blum K, Hummel D. Main-chain and side-chain crystallinities of isotactic poly(1-alkylethylene)s 1. Thermoanalysis and x-ray diffraction of the higher homologues. Makromol Chem 1976;177:108995. [147] Alamo RG, Viers BD, Mandelkern L. Phase structure of random ethylene copolymers: A study of counit content and molecular weight as independent variables. Macromolecules 1993;26:57407. [148] Jones JR, Liotta CL, Collard DM, Schiraldi DA. Crosslinking and modication of poly(ethylene terephthalate-co2,6-anthracenedicarboxylate) by diels-alder reactions with maleimides. Macromolecules 1999;32:578692. [149] Conner DM, Allen SD, Collard DM, Liotta CL, Schiraldi DA. Effect of linear comonomers on the rate of crystallization of copolyesters. J Appl Polym Sci 2001;80: 2696704. [150] Ma H, Vargas M, Collard DM, Kumar S, Schiraldi DA. Crosslinking studies on poly(ethylene terephthalate-co-1,4phenylene bisacrylate). J Appl Polym Sci 2004;91:1698702. [151] Rosu RF, Shanks SN, Bhattacharya N. Synthesis and characterization of branched poly(ethylene terephthalate). Polym Int 1997;42:26773.

[119] McKee MG, Wilkes GL, Colby RH, Long TE. Correlations of solution rheology with electrospun ber formation of linear and branched polyesters. Macromolecules 2004;37: 17607. [120] Juliani, Archer LA. Relaxation dynamics of entangled and unentangled multiarm polymer solutions: experiment. Macromolecules 2002;35:695360. [121] Gibson P, Gibson-Schreuder H, Rivin D. Transport properties of porous membranes based on electrospun bers. Colloids and surfaces A: physiochemical and engineering aspects 2001;187188:46981. [122] Reneker DH, Chun I. Nanometre diameter bers of polymer, produced by electrospinning. Nanotechnology 1996;7: 21623. [123] Deitzel JM, Kleinmeyer D, Harris D, Beck Tan TC. The effect of processing variables on the morphology of electrospun nanobers and textiles. Polymer 2001;42: 26172. [124] Taylor GI. In: Proceedings of the royal society of London, series A, 1969. p. 313 [see also p. 453]. [125] Reneker DH, Yarin AL, Fong H, Koombhongse S. Bending instability of electrically charged liquid jets of polymer solutions in electrospinning. J Appl Phys 2000;87:453147. [126] Lee KH, Kim HY, Bang HJ, Jung YH, Lee SG. The change of bead morphology on electrospun polystyrene bers. Polymer 2003;44:402934. [127] Hohman MM, Shin M, Rutledge G, Brenner MP. Electrospinning and electrically forced jets 1. Stability theory. Phys Fluids 2001;13:220120. [128] Shin YM, Hohman MM, Brenner MP, Rutledge GC. Experimental characterization of electrospinning: the electrically forced jet and instabilities. Polymer 2001;42: 995567. [129] Deitzel JM, Kosik W, McKnight SH, Beck Tan TC, Desimone JM, Crette S. Electrospinning of polymer nanobers with specic surface chemistry. Polymer 2002; 43:10259. [130] Grafe T, Graham K. Polymeric nanobers and nanobers webs: a new class of nonwovens. Nonwoven Technology Review 2003;515. [131] Kenawy E-R, Layman JM, Watkins JR, Bowlin GL, Matthews JA, Simpson DG, Wnek GE. Electrospinning of poly(ethylene-co-vinyl alcohol) bers. Biomaterials 2003; 24:90713. [132] Matthews JA, Wnek GW, Simpson DG, Bowlin GL. Electrospinning of collagen nanobers. Biomacromolecules 2002;3:2328. [133] Megelski S, Stephens JS, Chase DB, Rabolt JF. Micro and nanostructured surface morphology on electrospun polymer bers. Macromolecules 2002;35:845667. [134] Lee KH, Kim HY, La Y M, Lee DR, Sung NH. Inuence of mixing solvent with tetrahydrofuran and N,N-dimethylformamide on electrospun poly(vinyl chloride) nonwoven mats. J Polym Sci Part B: Polym Phys 2002;40:225968. [135] Fong H, Chun I, Reneker DH. Beaded nanobers formed during electrospinning. Polymer 1999;40:458592.

M.G. McKee et al. / Prog. Polym. Sci. 30 (2005) 507539 [152] Ramakrishnan S, Jayakannan M. Effect of branching on the crystallization kinetics of poly(ethylene terephthalate). J Appl Polym Sci 1999;74:5966. [153] Jabarin SA. Crystallization kinetics of polyethylene terephthalate II. Dynamic crystallization of PET. J Appl Polym Sci 1987;34:97. [154] Albertsson A-C, Varma IK. Aliphatic polyesters: synthesis, properties, and applications. Adv Polym Sci 2002;157:140. [155] John G, Tsuda S, Morita M. Synthesis and modication of new biodegradable copolymers: serine/glycolic acid based copolymers. J Polym Sci Part A: Polym Chem 1997;35: 19017. [156] Tasaka F, Miyazaki H, Ohya Y, Ouchi T. Synthesis of combtype biodegradable polylactide through depsipeptide-lactide copolymer containing serine residues. Macromolecules 1999;32:63869. [157] Kim M-N, Kim K-H, Jin H-H, Park J-K, Yoon J-S. Biodegradability of ethyl and n-octyl branched poly(ethylene adipate) and poly(butylene succinate). Eur Polym J 2001;37: 18437.

539

[158] Tasaka F, Ohya Y, Ouchi T. Synthesis of novel comb-type polylactide and its biodegradability. Macromolecules 2001; 34:5494500. [159] Weinhold A, Gurny R. Controlled and/or prolonged delivery of peptides from the hypothalamic pituitary axis. Eur J Pharm Biopharm 1997;43:11531. [160] Bodmer D, Kissel T, Traeschslin E. Factors inuencing the release of peptides and proteins from biodegradable parenteral depot systems. J Control Release 1992;21:12938. [161] Kissel T, Li YX, Volland C, Gorich S, Koneberg R. Parenteral protein delivery systems using biodegradable ABA-block copolymers. J Control Release 1996;39:31526. [162] Breitenbach A, Kissel T. Biodegradable comb polyesters: part 1 Synthesis, characterization, and structural analysis o poly(lactide) and poly(lactide-co-glycolide) grafted onto water-soluble poly(vinyl alcohol) as backbone. Polymer 1998;39:326171. [163] Breitenbach A, Li YX, Kissel T. Branched biodegradable polyesters for parenteral drug delivery systems. J Control Release 2000;64:16878.

Вам также может понравиться