Download as pdf or txt
Download as pdf or txt
You are on page 1of 102

Version of Record: https://www.sciencedirect.

com/science/article/pii/S0079670018303502
Manuscript_5c6d8e270c341a3cee91eb2591f186ff

Recent Advances in Thermoplastic Elastomers from Living Polymerizations: Macromolecular

Architectures and Supramolecular Chemistry

Weiyu Wang1,2,3 , Wei Lu1, Andrew Goodwin1, Huiqun Wang1, Panchao Yin2, Nam-Goo Kang1*, Kunlun
Hong3*, Jimmy W. Mays1*
1. Department of Chemistry, University of Tennessee, Knoxville, TN 37996, USA
2. South China Advanced Institute for Soft Matter Science and Technology, South China University of
Technology, Guangzhou 510640, P. R. China
3. Center for Nanophase Materials Sciences, Oak Ridge National Laboratory, Oak Ridge, TN 37831, USA

Abstract
Thermoplastic elastomers (TPEs) have found use in a wide range of applications, such as adhesives,
elastomers, coatings, fibers, and in additive manufacturing techniques such as 3D printing. Despite their
omnipresence, the need for advanced TPEs with adaptive properties is continuously growing. Along with a
brief historical introduction, this review presents an overview of typical structure-property relationships for
various TPEs and discusses the design principles of TPEs from a synthetic chemistry perspective. Recent
advances in TPEs with different macromolecular architectures, including linear ABA triblock copolymers,
ABC triblock terpolymers, multiblock copolymers, star copolymers, graft copolymers, bottlebrush
polymers, and hyperbranched polymers are reviewed. Service temperatures and mechanical properties of
the different materials are compared in each section. Incorporating various supramolecular interactions into
different macromolecular architectures as a means to further extend the range of TPE applications is also
discussed. Future opportunities for TPE research in both academia and industry are addressed as
perspectives.

Key Words
Thermoplastic elastomer, living/controlled polymerization, macromolecular architecture, mechanical
properties, glass transition temperature, supramolecular chemistry

* Corresponding Authors: nkang1@utk.edu, hongkq@ornl.gov, jimmymays@utk.edu,

© 2019 published by Elsevier. This manuscript is made available under the Elsevier user license
https://www.elsevier.com/open-access/userlicense/1.0/
Contents

1. Introduction
2. Synthetic design principles
2.1. The choice of the monomer
2.2. The composition of the polymer
2.3. Macromolecular architectures
3. Linear polymers for TPEs
3.1. ABA triblock copolymer type TPEs
3.2. ABC triblock terpolymers type TPEs
3.3. Multiblock polymer type TPEs
4. Branched copolymers for TPEs
4.1. Star branched copolymers for TPEs
4.2. Graft copolymers for TPEs
4.3. Bottlebrush Copolymers for TPEs
4.4. Copolymers with other branched architectures
5. TPEs based on supramolecular chemistry
5.1 TPEs with hydrogen bonding
5.2 TPEs with ionic interactions
5.3 TPEs with π-π stacking
5.4 TPEs with metal-ligand coordination.
6. Conclusions and Perspectives
Nomenclature
ϕbridge bridge fractions PpCS Poly(p-chlorostyrene)
χ thermodynamic parameter PDI molecular weight distribution
Ade- Adenine PDL Poly(ε-decalactone)
ATRP atom transfer radical polymerization PDLA poly(D-lactide)
Ba poly(n-butyl acrylate)-co- PDMA poly(dodecyl methacrylate)
polyacrylamide PDMS polydimethylsiloxane
nBA n-butyl acrylate PE Polyethylene
tBA tert-butyl acrylate PEB poly(ethylene-co-butylene)
BF Benzofulvene PEMA poly(ethyl methacrylate)
bBr-iBE 1,2-bis(bromoisobutyryloxy) ethane PEO Poly(ethylene oxide)
BTBTMB 1,4-bis(thiobenzoylthiomethyl)benzene PEP poly(ethylene-co-propylene)
CHD 1,3-cyclohexadiene PID Polyindene
CHVE cyclohexyl vinyl ether PLLA poly(L-lactide)
CL Caprolactone PMBL poly(α-methylene-γ-butyrolactone)
COPAs polyamide-based thermoplastic PMCH poly(1,3-methylenecyclohexane)
elastomers PMCL poly(6-methyl-ε-caprolactone)
COPEs polyether ester based thermoplastic PMD Polymenthide
elastomers PMMA polymethacrylate
CRP controlled radical polymerization PODMA poly(octadecyl methacrylate)
CSA chain shuttling agent PP Polypropylene
CTA chain transfer agent a-PP atactic-polypropylene
DEMDBA diethyl meso-2,5-dibromoadipate i-PP isotactic-polypropylene
DPE 1,1-diphenylethylene s-PP syndiotactic-polypropylene
DPNR deproteinized natural rubber PS Polystyrene
DL ε-decalactone PSM poly[styrene-alt-(maleic anhydride)]
G’ storage modulus PTCDVE poly(tricyclodecane vinyl ether)
HA 3-hydroxybutyrate PVA poly(vinyl alcohol)
HPMC hydroxypropyl methylcellulose PVBA N-isopropyl-4-vinylbenzylamine
LAP living anionic polymerization P2VP poly(2-vinyl pyridine)
LA system Lewis base and di-phenoxyalkyl RAFT reversible addition-fragmentation chain
aluminum transfer polymerization
MAM polymethacrylate-b-polyacrylate-b- ROMP ring opening metathesis polymerization
MCL 6-methyl-ε-caprolactone [RS]-PHB Poly([RS]-3-hydroxybutyrate)
MD Menthide SBCs styrenic block copolymers
Me entanglement molecular weight SBS polystyrene-polybutadiene-polystyrene
MGC multigraft copolymers SCFT self-consistent field theory
MMA methyl methacrylate SG1 N-tertbutyl-1-diethylphosphono-2,2-
βMδVL β-methyl-δ-valerolactone dimethylpropyl nitroxide
NBMP 5-norbornene-2-methylpropanoate SIP polystyrene-b-polyisoprene-b-poly(2-
NMP nitroxide-mediated radical vinyl pyridine
polymerization SIS polystyrene-polyisoprene-polystyrene
PACP polyacenaphthylene Tg glass transition temperature
PADVE poly(2-adamantyl vinyl ether Thy- Thymine
PAdMA poly(1-adamantyl methacrylate Tm melting temperature
PpAdSt poly(p-adamantyl styrene) TPEs thermoplastic elastomers
PiB Polyisobutylene TPOs rubbery-polyolefin blends
PiBMA poly(isobornyl methacrylate) TPUs thermoplastic polyurethanes
PtBMA poly(tert-butyl methacrylate) TPVs dynamically vulcanized polymer blends
PtBVE poly(tert-butyl vinyl ether) UPy 2-ureido-4[1H]-pyrimidinone
PCHE poly(cyclohexylethylene) UST upper service temperature
PCL poly(ε-caprolactone)
4

1. Introduction

Consisting of an elastic matrix that is physically crosslinked by plastic domains, thermoplastic

elastomers (TPEs) are rubbery materials that are suitable for use with plastic processing techniques

[1, 2]. Using high throughput techniques, such as melt extrusion and injecting molding, TPE

products can be manufactured in large volumes with short production cycles [3, 4]. Based on

chemical composition and morphology, there are six groups of commercially available TPEs [3]:

(1) styrenic block copolymers (SBCs), (2) thermoplastic polyurethanes (TPUs), (3) polyamide-

based thermoplastic elastomers (COPAs), (4) polyetherester-based thermoplastic elastomers

(COPEs), (5) rubbery-polyolefin blends (TPOs), and (6) dynamically vulcanized polymer blends

(TPVs). Among these, the first four groups are normally composed of block copolymers, whereas

TPOs and TPVs are mainly polymer blends. These TPEs materials have been widely applied in

fields including automotive, footwear, medical devices, sportings goods, etc. [1, 5-7].

As a brief historical introduction to thermoplastic elastomers, polyurethanes fibers started to

emerge in the plastic/rubber market in the early 1950s with the discovery and development of basic

diisocyanate addition [8]. In 1952, Snyder first patented linear copolyester fibers with stress and

strain characteristics comparable to those of natural rubber. These materials are considered as

multiblock TPUs [9]. Many other polyurethane-based thermoplastic gum elastomers were

introduced by Bayer, Dupont, and Goodyear in the late 1950s [10]. Polyurethane-based TPEs

showing high abrasion resistance with excellent elasticity and tensile strength were

commercialized by B.F Goodrich in the 1960s [11].

A milestone in the development of TPEs occurred in 1965 was when Milkovich and Holden

from Shell Chemical Company (USA) synthesized (by living anionic polymerization) and

characterized ABA triblock copolymers based on polystyrene (PS) as the hard segment (A block)
5

and polyisoprene (PI) as the rubbery matrix (B block) [12]. The use of the termination free (also

termed “living”) anionic polymerization techniques allows styrenic TPEs (SBCs), such as

polystyrene-polybutadiene-polystyrene (SBS) or polystyrene-polyisoprene-polystyrene (SIS), to

be prepared with predicted molecular weight, narrow molecular weight distribution and

quantitative yield, containing negligible impurities, in a short production cycle [13]. The stress-

strain and elastic recovery of SBCs display similar behavior to conventional vulcanized rubbers

without filler reinforcement or crosslinking [14]. This “new type” of TPE with lower cost and

well-defined structure, along with the development of living anionic polymerization [15, 16],

stimulated numerous research activities in polymer thermodynamics [17], self-consistent field

theory (SCFT) [18], and solution properties [19] of these materials in both academic research and

industrial applications. Shell Chemical Company quickly realized the potential of processing

styrenic TPEs (later given the trade name as Kraton®) with injection molding and melt extrusion.

After more than 50 years, SBCs have developed into one of the major products in the polymer

industry, with applications spanning the range from items in our daily life (soft touch toothbrush

grips) to use in the construction industry, in medical devices, and in many other advanced systems

[20, 21]. According to a report, the global market size of TPEs was over USD 12 billion in 2015

and is estimated to be worth over USD 20 billion by 2023 [22].

Besides SBCs, many other types of TPEs were developed from the 1960s to 1980s. Hercules

Inc. patented the first TPE based on mixtures of elastic poly(ethylene-co-propylene) (PEP) with

30 to 70% weight percentage of polyethylene (PE) in 1966 [23] . In the late 1970s, Monsanto

Company developed a dynamic vulcanization process to chemically crosslink blends of PEP-diene

and PP. The product was commercialized under the trade name of Santoprene® (a product of TPV)

[24]. DuPont commercialized a type of copolyester elastomer (COPE), with the tradename of
6

Hytrel®, that combined good mechanical properties with the chemical and heat resistance of certain

polyesters [25]. In 1982, Dow Chemical Company developed Estamid®, a segmented COPA with

low density, superior mechanical properties at low temperature (-40 °C) and a higher upper service

temperature above 200 °C [26].

During the 1990s, many companies strategized their TPE research focus on specific market

applications by adding new functionalities and aiming to harness sustainable resources to improve

properties of existing TPE systems. The toolbox of synthetic polymer chemistry has been boosted

tremendously by developments made in living/controlled polymerization techniques, including

living anionic polymerization [15, 27], cationic polymerization [28], ring opening metathesis

polymerization (ROMP) [29], atomic transfer radical polymerization (ATRP) [30], reversible

addition-fragmentation chain transfer (RAFT) polymerization [31], and nitroxide-mediated radical

polymerization (NMP) [32]. Well defined polymers with complex macromolecular architectures

and various functionalities have been prepared for application in the field of TPEs. Two successful

examples of transferring advancements in polymerization to TPE products are the all acrylic TPEs:

Nanostrength® from Arkema [33] and KurarayTM from Kuraray [34].

2. Synthetic design principles

When designing a polymer for TPE applications, three parameters affect the structure-property

relationships of the elastomer: (1) the choice of monomer pairs, (2) the composition of the polymer,

and (3) the macromolecular architecture. It should be noted that thermoplastic elastomers are rarely

used in their pure chemical form for commercial applications, however the topic of the formulation

process is beyond the scope of this review. The design principles discussed here are mainly from
7

the synthetic chemistry perspective and the reported physical properties are based solely on the

polymeric material in the absence of any compounding and processing additives.

2.1 The choice of monomers

The basic building blocks of TPE materials are elastic components that form a rubbery matrix

and glassy components that form plastic domains. The choice of the monomer pairs directly

decides the glass transition temperature (Tg) range, which is associated with TPEs' service

temperature range. When the service temperature is lower than the Tg of the rubbery matrix, TPEs

act as stiff, brittle plastics and display a high storage modulus (G’). When the service temperature

is higher than the Tg or melting temperature (Tm) of the hard phase, TPEs become weak viscous

fluid materials. Between these transition temperatures lies the proper service temperature range of

TPEs where G’ reaches the rubbery plateau and the materials display reversible elongation

behavior upon application of moderate stresses (Figure 1) [35].

Figure 1. Dynamic mechanical analysis (DMA) response of a typical TPE with temperature

ramp/frequency sweep [35]. Copyright 1992, Reproduced with permission from the American

Chemical Society.
8

In addition to regulating the service temperature range, the choice of monomer pairs also dictates

the thermodynamic parameter, χ, the key parameter that describes the strength of phase separation

within block copolymers [36, 37]. For different block copolymers with the same composition,

number of repeat units (N) and macromolecular architecture, the higher the value of χ the stronger

the phase separation [38]. When using polymers as TPEs, the materials display thermoplastic

elastomer behavior usually when the rubbery matrix is physically crosslinked by the glassy

domains. The benefit of choosing monomer pairs with high χ values is that the block polymers

form stronger physical crosslinks at relatively low molecular weight [38]. Generally, the product

of χ and N needs to be higher than 10.5 to get phase separation for diblock copolymers; whereas

for multiblock copolymers, the prerequisite value is even higher. A product value higher than 60

indicates a strong segregation and is thus preferred for the material design. When choosing two

monomers with lower χ, phase blending between two blocks occurs at moderate molecular weight.

For example, polystyrene derivatives with bulky lipophilic pendent groups, such as tert-butyl- or

adamantyl-, at the para- position might cause phase blending with polydienes [39, 40]. Thus, in

order to achieve similar mechanical properties as SBS, the overall molecular weight of poly(tert-

butyl styrene) containing TPE needs to be higher than that of its SBS counter-part. Therefore, the

choice of monomer impacts not only the phase separation but also their mechanical performance.

Notice from Figure 2, poly(styrene sulfonate) has the largest estimated χ with respect to

polystyrene [41]. Post chemical modifications such as sulfonation [42] or hydrogenation [43] may

increase the interaction parameter between two polymers, driving the system to stronger phase

separation, and may enhance the mechanical properties as well. For example, by hydrogenating PI

into poly(ethylene-co-propylene) (PEP), the temperature (T) dependence of χ for PS and PI is χPS-
9

PI = -0.0258 + 37.9/T whereas for PS and PEP is χPS-PEP = 0.050 + 8.00/T [43]. When T = 373K,

the value of χPS-PEP is larger than χPS-PI.

Figure 2. Estimated χ values of various polymers with respect to polystyrene [41]. Copyright

2012, Reproduced with permission from the American Association for the Advancement of

Science.

2.2 The composition of the polymer

The most well-studied TPEs are SIS and SBS triblock copolymers. Based on experimental data,

Morton specified the proper composition of ABA triblock copolymers for TPE applications: the

Mn of PS was targeted at 10 to 20 kg/mol (or 13 to 33 vol%), whereas the Mn of polydiene was

targeted at 40 to 80 kg/mol (or 67 to 87 vol%) [44]. In order to evaluate the effects of overall

molecular weight to the mechanical properties, Holden prepared three SBS triblock copolymers

with the same composition (around 27 wt% of PS) but different overall molecular weight of 73

kg/mol, 102 kg/mol and 150 kg/mol [45]. All samples displayed the same extension at break and

ultimate tensile strength. However, as the molecular weight increased, the viscosity increased

significantly at low shear rate. For SBS with similar overall molecular weight but different PS

weight percentage (wt%), the elongation at break decreased as the wt% of PS increasing (Figure

3.a). With 13 % of PS, no strain hardening behavior was observed. Even with an elongation at

break around 1000%, the overall mechanical properties of the sample were limited. As the wt% of
10

PS increased from 13% to 28% and 39%, ultimate tensile strength increased, and strain hardening

was observed. Samples started to display typical TPE stress-strain behaviors. Once a volume

fraction of PS reached 53% or above, the formation of lamellar phase disrupted the continuous

rubbery matrix that led to an irreversible yield point at very low elongations. The hardness and the

initial modulus increased with the wt% of PS [45].

Figure 3. a) Illustration of stress-strain plots for SBS [45]. Copyright 1969, Reproduced with

permission from Wiley Periodicals, Inc. b) Phase diagram of PS-PI diblock copolymer near

order-disorder transition [46]. Copyright 1995, Reproduced with permission from the American

Chemical Society.

In the last two decades of the 20th century, advances in transmission electron microscopy

(TEM), small-angle X-ray scattering (SAXS), and self-consistent field theory (SCFT) significantly

facilitated the fundamental understanding of the structure-property relationships of polymers.

Figure 3.b shows the composition dependence of the morphology of SI diblock copolymers

(Figure 3.b) [46]. The understanding of phase separation behavior of block copolymers connected

the composition with morphology and translated the mechanical properties into the structure-

property relationship of TPEs [46, 47]. In Holden’s design, PS and PB in all samples were above
11

strong phase separation limit where distinct domains of PS and PB were formed. The minor

component of PS microphase separated from polydiene and formed either spherical (less than 20

vol% PS) or cylindrical (20 vol% to 35 vol% of PS) glassy domains, which acted as physical

crosslinks and reinforced the entangled polydiene rubbery matrix. In summary, when designing an

ABA triblock copolymer type TPEs, three aspects need to be taken into consideration: (1) under

the designed composition, the B block needs to form a continuous rubbery matrix to provide

enough elasticity. (2) Overall molecular weight needs to be high enough to drive micro-phase

separation for efficient stress reinforcement. (3) Molecular weight should not be too high

considering that high viscosity may cause difficulty in processing [45]. The chemical structures of

most glassy blocks and the elastic blocks, as well as their glass transition temperatures or melting

temperatures, have been summarized in Table 1 for hard blocks and Table 2 for elastic blocks.

Table 1. Summary of chemical structures and Tg (Tm) of hard blocks

Tg 65 ~ 95 oC 180 oC 110 oC 123 oC 182 oC 130 oC 180 oC


ref 44, 45, 116,117, 168, 169, 48, 118 122 122,257 49 39, 117 40, 50
172, 173,202-204, 235, 243, - 121
248-251, 262, 276, 280, 282

Tg 120 ~ 135 oC 90 oC 116 oC 160 oC 133 oC 118 oC


ref 33, 34, 66, 69-71, 83- 71 71 72-74 61, 75, 76 59, 60 81
85, 122, 202, 203, 277,
278, 281
12

Tg 140 oC 145 ~160 oC 225 oC 250 oC 88 oC 150 oC 77 oC


ref 106, 107 104,105,108,109 117,123 125 129, 131 133, 134 132

Tg 130 oC 134 oC 60 oC 147 oC 200 oC


ref 177, 183 184-187 188 115, 218 107

Tg 60 ~ 175 oC 170 ~ 190 oC 105 oC 150 oC 130 ~ 149 oC


ref 141-149, 220 163, 164 261 40 69,70

Table 2. Summary of chemical structures and Tg of elastic blocks

Tg -64 oC -90 oC -39 oC -52 oC


ref 48, 109, 141,239, 248- 44, 45, 66, 69-71, 73, 49 143
251, 270-275, 280 74, 80, 81, 106, 107

Tg -40 oC -21 oC -35 oC -50 oC -52 oC


ref 76, 83-90, 259-262, 86, 88 86, 88 83, 85, 86 132-134
264, 276-279

Tg -58 oC -43 oC -52 oC


ref 156 155 150, 151

Tg -30 oC -63 oC -25 oC


ref 153, 154 153 147-149
13

Tg -61 oC -20 oC -23 oC


ref 116 – 128, 142, 254-258 145 332
2.3 Macromolecular architectures

To ensure the formation of the physical crosslinking junctions necessary to reinforce the

mechanical strength of the materials, both ends of the elastomer segment need to be linked to

glassy domains [45]. For example, in ABA triblock copolymers, A blocks are the hard block and

B block is the elastic block. AB diblock copolymers and BAB triblock copolymers (with the

elastomeric domain as the end blocks) do not show the characteristic behavior of TPEs - reversible

elongation. However, AB and BAB type copolymers may display TPE characteristics if they are

endowed with appropriate supramolecular interactions (discussed in detail in Section 5). Other

linear type polymers, such as ABC triblock terpolymer or AB multiblock copolymers, could also

serve as TPEs with proper design of the composition and choice of monomers. One benefit of

designing TPEs with these latter two macromolecular architectures is to suppress the formation of

loops and enhance the formation of bridges in the elastic domains. Details of these polymers are

discussed in Section 3.2 ABC triblock terpolymer type TPEs and Section 3.3 multiblock polymer

type TPEs.

In contrast to linear polymers, branched polymers provide the opportunity to decouple the

morphological dependence on composition, seen for linear block copolymers, by choice of

different branched macromolecular architectures, such as miktoarm star copolymers or graft

copolymers. The asymmetry incorporated at the junction points between two components provides

opportunities to prepare thermoplastic elastomers containing more than 50% of hard blocks.

Section 4 (branched polymers for TPEs) in this review is mainly focused on elucidating structure-

property relationships of TPEs based on branched polymers.


14

As a summary of the synthetic design principles of TPEs (Figure 4): the choice of monomers

directly determines the service temperature range and the strength of the phase separation, which

all contribute to the mechanical properties. To be applied as TPEs, a self-reinforced morphology

with a continuous rubbery matrix and discontinuous dispersed plastic domains, such as those

existing with spherical and cylindrical morphologies, is critical. The morphology depends on both

composition and macromolecular architecture. All these above-mentioned parameters ultimately

govern the mechanical properties of the designed materials.

Figure 4. Summary of synthetic design principles

3. Linear polymers for TPEs

As illustrated in Figure 5, most of the linear copolymers used as TPEs are ABA triblock

copolymers, ABC triblock terpolymers and (AB)n multiblock copolymers with A or C as the glassy

blocks and B as the elastic blocks. These polymers may be prepared by living anionic/cationic

polymerizations, ring opening polymerizations, controlled radical polymerizations, and living

coordination polymerizations. Section 3 discusses advances in preparing TPEs with linear

macromolecular architectures by using different polymerization methods or adopting a

combination of them.
15

Figure 5. Linear polymer architectures for TPE applications.

3.1. ABA triblock copolymer type TPEs

3.1.1. Polymers prepared by living anionic polymerization

Anionic polymerization can produce well-defined block copolymers in short manufacturing

times, with quantitative conversion, and in large volume. In industry, this termination free

polymerization technique is mainly applied to prepare styrenic block copolymers for various

applications. However, one issue associated with polystyrene (PS) containing materials is that as

the service temperature approaching the glass transition temperature of PS, the physically

crosslinked PS domains soften and flow, and the mechanical properties of the material drastically

decreases. Since the discovery of SBS and SIS triblock copolymer TPEs in the 1960s, tremendous

efforts have focused on increasing the upper service temperature (UST) of SBCs without

significantly altering their production procedure [14].

3.1.1.1. TPEs based on polystyrene derivatives

To increase the upper service temperature, the first strategy is to use polystyrene derivatives

with a higher glass transition temperature such as poly(α-methyl styrene) (Tg ~173 °C) [48],

poly(α-methyl p-methyl styrene) (Tg ~183 °C) [49], poly(tert-butyl styrene) (Tg ~130 °C) [39], and

poly(p-adamantyl styrene) (PpAdSt, Tg ~ 203 °C) [40, 50]. For anionic polymerization of α-methyl

styrene and its derivatives, a polymerization temperature below its ceiling temperature is required

to suppress the depolymerization process caused by the enthalpic penalty imparted by the α-methyl
16

group [51]. To achieve a quantitative yield, the polymerization is thus generally performed at -78

°C in a polar solvent such as THF [52]. In contrast, the anionic polymerization of tert-butyl styrene

and adamantyl styrene may be carried out in hydrocarbon solvents at moderate temperatures, with

the advantage of reduced cost. However, the corresponding polymers have relatively low χ with

respect to polydienes. Thus, higher overall molecular weight is required to increase the strength of

phase separation, which again can lead to high viscosity during processing. Through catalytic

hydrogenation of polyisoprene into poly(ethylene-alt-propylene), the mechanical properties are

enhanced due to the stronger phase separation between PpAdSt and poly(ethylene-alt-propylene)

[40]. It is worth mentioning that incorporating the adamantyl group can significantly increase the

Tg of the polymer [53]. Many adamantyl containing polymers with styrene [54-56], butadiene [57,

58], methacrylate [59, 60] or acrylate [61] skeletons have great potential to serve as building blocks

for high-temperature TPEs with proper structural design or post-chemical modification.

Another classical approach in anionic polymerization to increase the glass transition temperature

of PS is to copolymerize styrene with 1,1-diphenylethylene (DPE), a bulky monomer incapable of

forming homopolymer due to inherent steric hindrance. The copolymerization behavior of DPE

with various anionic polymerizable monomers, such as isoprene, butadiene, and styrene and its

derivatives has been investigated by Yuki, Bates, and Hutchings [62-65]. Hutchings prepared a

perfectly alternative styrene/DPE alternative copolymer with a Tg around 180°C [64]. By

sequentially adding butadiene into the living anion of PS/DPE alternative copolymer, diblock

copolymers with narrow polydispersity and clear microphase separation were prepared [65]. Since

this polymerization process used benzene as the polymerization solvent, the combination of

styrene and DPE as the glass block, and butadiene or isoprene as the elastic block potentially offers
17

an industrially scalable process with which to prepare high temperature TPEs with tunable upper

service temperature.

3.1.1.2. TPEs based on poly(methyl methacrylate) derivatives.

The second anionic polymerization strategy to improve UST is to use methyl methacrylate

(MMA) or its derivatives as the hard segment. By using lithium chloride or other chelating agents

as an additive during polymerization, methyl methacrylate may be anionically polymerized in THF

at -78 oC. The PMMA synthesized by this method contains 75 % to 79 % of syndiotactic repeating

units and exhibits a glass transition temperature around 125 °C [66]. When mixing isotactic-

PMMA and syndiotactic-PMMA, a stereocomplex is formed which has a melting temperature

above 170 °C [67, 68]. The melting temperature is adjustable by using different annealing solvents

or different iso/syndio mixing ratios. The formation of semi-crystalline stereocomplexes reinforces

the aggregation of hard segments [69, 70]. Beside tacticity, the length of the alkyl substituent also

alters the glass transition temperature of poly(alkyl methacrylate)s [66, 71]. A series of

poly(meth)acrylates with different Tgs, such as poly(isobornyl methacrylate) (PiBMA, Tg ~ 202

°C) [72-74], poly(1-adamantyl methacrylate) (PAdMA, Tg ~ 202 °C ) [59, 60], and poly(1-

adamantyl acrylate) (Tg ~ 133 °C) [61, 75, 76], provides a versatile toolbox with which to tune the

mechanical performance and UST.

Conventionally, PMMA hard blocks are incorporated on both ends of polybutadiene. However,

to prepare elastic PB with high cis-1,4 microstructure, a hydrocarbon solvent is required for the

anionic polymerization [77-79]. Thus, an α,ω- difunctional polybutadiene living anion needs to be

prepared first in a hydrocarbon solvent. THF is then added into the living polymer solution,

followed by reducing the temperature to -78 oC and initiating MMA at the same temperature [69-

71, 73,74, 80, 81]. Other methacrylate derivatives, such as poly(ethyl methacrylate) (PEMA, Tg ~
18

90 °C), poly(tert-butyl methacrylate) (PtBMA, Tg ~ 116 °C) and PiBMA, can also be anionically

polymerized by similar methods [71]. The resulting triblock copolymers with polybutadiene as the

rubbery matrix displayed high ultimate stress at break and strain at break over 500 %. Notably,

when PiBMA was used as the rigid block, triblock copolymers exhibited 600 % strain at break

with ultimate tensile stress of 2.2 MPa even when the temperature reached 150 °C. Tuning UST

from 130 °C to 149 °C was possible by randomly copolymerizing iBMA with MMA using

different feed ratios [74]. UST of styrenic TPEs could also be enhanced by incorporating MMA

as the minor component (13 wt%) in linear PMMA-PS-PB-PS-PMMA pentablock terpolymers

TPEs (32 MPa ultimate stress, 900 % strain at break) [81].

3.1.1.3. All-acrylic TPEs

The above mentioned TPE systems contain polybutadiene as the elastic block. In all-acrylic

TPEs, poly(alkyl acrylate) is used as the soft block, resulting in better UV and oxidation resistance

due to the lack of unsaturated bonds [82-90]. Additionally, all-acrylic TPEs have many advantages

over SBCs including optical transparency, versatility of adhesion, weatherability, oil resistance,

printability, compatibility with fillers, abrasion resistance and low viscosity [34, 91].

The living anionic polymerization of acrylates, especially with primary and secondary carbon

linked to the ester group, is challenging due to intrinsic side reactions including backbiting

reactions of propagating enolate anions and aggregation of active chain ends [92]. Although many

attempts have been made, few acrylate monomers have been successfully polymerized anionically

with low PDI and predicted molecular weights. The successful attempts employed monomers with

bulky pendant groups to prevent or limit the side reactions [61, 93, 94]. Transalcoholysis of tert-

butanol to n-butanol bypassed the difficulties in the direct anionic polymerization of n-butyl

acrylate (nBA). Tong reported the synthesis of PMMA-b-PtBA-b-PMMA by sequential anionic


19

polymerization of MMA and tert-butyl acrylate (tBA) using the sec-butyllithium/1,1-

diphenylethylene/lithium chloride (sec-BuLi/DPE/LiCl) initiation system. The polymerization

was performed in THF at -78 oC. The transalcoholysis was performed by adding excess n-butanol

in the presence of p-toluenesulfonic acid [95]. By keeping the molecular weight of PnBA at 100

kg/mol and varying the weight percentage of PMMA from 9.1 % to 50 %, a series of PMMA-b-

PnBA-b-PMMA all acrylic TPEs was prepared. As the wt% of PMMA increased, the ultimate

tensile strength was increased from 1.8 MPa to 16.1 MPa where the elongation at break decreased

from 1016 % to 228 % [83]. The morphology evolved from spheres, to cylinders, to lamellae [95].

The mechanical properties of all-acrylic TPEs are generally inferior to styrenic TPEs because

of the high entanglement molecular weight (Me) of poly(alkyl acrylate) (Me for PB is 1,700 g/mol

whereas Me for PnBA is 28,000 g/mol) and the relative low χ between PMMA and PnBA [89] as

compared to PS and polydienes. Larger Me leads to fewer entanglements in the elastic matrix.

Therefore, the mechanical strength and strain at break tend to be lower. The ultimate tensile

strength increases with a decrease of the Me of the elastic block (Figure 6). By altering the elastic

block of all-acrylic TPEs from ethyl, n-propyl, n-butyl, to iso-octyl acrylate, the ultimate strength

of resulting TPEs decreased linearly with respect to Me-1 [86]. The Me of each polyacrylate elastic

block is increased due to higher packing lengths of the polymer chain [96]. With similar

composition (22 wt% of PMMA) and overall molecular weight, all acrylic TPEs with poly (iso-

octyl acrylate) [87] as the middle block showed the highest ultimate stress (16.2 MPa) but the

lowest strain at break (390 %) as compared to TPEs with other middle blocks [89].
20

Figure 6. Ultimate tensile strength as a function of 1/Me. In this figure M is PMMA; I is poly

(iso-octyl acrylate); nB is poly(n-butyl acrylate); nP is poly(n-propyl acrylate); E is poly(ethyl

acrylate); S is PS; IB is poly(iso-butylene); IP is polyisoprene [86]. Copyright2000, Reproduced

with permission from the American Chemical Society.

The effects of Me on elongation at break are also seen by comparing different TPEs composed

of PS or PMMA as the hard blocks, and PB or PnBA as the soft block. The elongation at break of

SBS is generally around 1000% [97]. By changing the hard block from PS to PMMA, the

elongation at break of PMMA-b-PB-b-PMMA could reach 1000% as well [80]. However, by

changing the soft blocks from PB to PnBA, the elongation at break decreased significantly. With

a similar molecular weight to SBS, the elongation at break of PS-b-PnBA-b-PS is less than 600%

[98], and the elongation at break of PMMA-b-PnBA-b-PMMA was around 500% [89]. The

average chain entanglement molecular weights for above-mentioned polymers are summarized in

Table 3.
21

Table 3. Average chain entanglement molecular weight (Me) for polymers [89]. Copyright 2001,

Reproduced with permission Elsevier Science Ltd.

Polymer M × 10 (g/mol)

Polyisoprene 6.1

Polybutadiene 1.7

Poly(ethylacryate) 11

Poly(n-propylacrylate) 16

Poly(n-butylacrylate) 28

Poly(isooctylacrylate) 59

Despite the success in synthesis of all acrylic TPEs and establishment of structure-property

relationships, the above-mentioned approach has two major obstacles for industrialization: (1) low

polymerization temperature, generally -78 oC and (2) post-polymerization modification, such as

transesterification. Kuraray Co. Ltd. developed a new anionic polymerization initiation system

using a mixture of a Lewis base and di-phenoxyalkyl aluminum (LA system, Figure 7) [99]. The

use of the robust alkyl aluminum additive allowed the reaction to be performed at -10 oC for the

alkyl acrylate monomer, and 50 oC for the alkyl methacrylate monomer. Direct anionic

polymerization of nBA was also achieved [99, 100]. This two-step monomer feed, living

polymerization method stabilized the enolate anion and resulted exclusively in polymethacrylate-

b-polyacrylate-b-polymethacrylate (MAM) triblock copolymers having narrow molecular weight

distribution (PDI = ~ 1.2) without residual homo- or diblock chains [85, 100-102]. The relatively

mild temperature considerably decreased the energy consumption and created opportunities for

successful commercialization of all acrylic TPEs [34, 99].


22

Figure 7. The stabilization structure of the LA system [99]. Copyright 2015, Reproduced with

permission from Springer Nature.

3.1.1.4 TPEs based on rigid polydienes.

The synthetic strategy adopted to increase the UST of TPEs was to replace PS with other higher

Tg polymers. Beside styrene and methacrylate derivatives, 1,3-cyclohexadiene (CHD) [103] and

benzofulvene (BF) [104, 105] are two conjugated monomers that can be anionically polymerized

in hydrocarbon solvents at mild temperatures.

The anionic polymerization of CHD suffered from side reactions due to the weakly acidic proton

on the monomer, leading to homopolymer and diblock copolymer during the triblock

copolymerization. For PCHD-PB-PCHD triblock copolymer with 30% of PCHD, the polymer

exhibited 10.2 MPa ultimate stress and 290% strain at break [106]. After partial hydrogenation of

PB, the resulting polymer exhibits 14.0 MPa ultimate stress and 570% strain at break. Noted that,

complete hydrogenation of PCHD results in a polyolefin, poly(cyclohexylene), with Tg above

231°C [107].

Polybenzofulvene is another high-temperature polymer with the capacity to copolymerize with

isoprene in both benzene [105, 108] and THF [104]. PBF-b-PI-b-PBF triblock copolymer with 14

vol% of PBF (labeled as FIF-14 in Figure 8) displayed 14.3 MPa ultimate stress and 1390 % strain
23

at break, which is competitive with Kraton D1112P (Figure 8) [109]. Similar to elastomeric

polydienes, the microstructure of PCHD and PBF depends on the initiator, additive employed, and

polymerization solvent. The glass transition temperature of polydienes normally increases with the

percentage of 1, 2- microstructure. Thus, incorporating polydiene as the hard block offers

opportunities to prepare TPEs with tunable service temperature while using the same feeding

monomers.

Figure 8. Dynamic mechanical analysis and stress-strain curves for PBF-PI-PBF [109].

Copyright 2016, Reproduced with permission from the American Chemical Society.

3.1.1.5. TPEs by hydrogenation

Polydiene containing SBCs suffer from two major disadvantages: (1) Their limited upper

service temperature prevents many advanced applications of SBCs. (2) The polydiene middle

blocks is susceptible to degradation on exposure to strong UV light and oxidation. Parallel to the

efforts on investigating alternative monomers, polymer chemists have also adapted various post-

chemical modifications, especially hydrogenation, to enhance the mechanical properties and

chemical resistance of SBCs. To overcome the oxidative vulnerability of the unsaturated

polydienes in SBCs, hydrogenation of PB into poly(ethylene-co-butylene) (PEB), or PI into


24

poly(ethylene-co-propylene) (PEP), has been widely applied in TPE applications. The resulting

SBCs are known as SEBS or SEPS [110]. This process improved the structural integrity of CBCs

at higher temperature, as well as their resistance to UV and oxidation. Compared with SBS, the

elongation of SEBS is reduced since the Me of PEB is larger than that of PB. The improved strength

and significantly enhanced modulus are attributed to the limited interfacial volume due to the

higher χ between PEB and PS [110]. The mechanical property benefits achieved through

hydrogenation are also evidenced by TPEs containing PpAdSt [40], PCHD [107], polyindene

[111] and poly(adamantyl butadiene) [57].

Another catalytic hydrogenation approach is to hydrogenate PS into poly(cyclohexylethylene)

(PCHE, Tg ~ 147 °C), producing TPEs with higher UST and better thermal stabilities [112, 113].

Originally, PCHE was developed as an alternative optical storage material to polycarbonate. The

complete hydrogenation of PS reduces the density from 1.06 g/cm3 to 0.947 g/cm3, which, by

increasing the volume of product, offsets the cost of hydrogenation. However, the hydrogenation

also increases the Me of PS from 13,309 g/mol to 40,000 g/mol, resulting in a brittle plastic [112].

In order to increase the toughness of PCHE, anionic synthesized SBS triblock and SBSBS

pentablock copolymers were hydrogenated into CEC and CECEC block copolymers where C

represents PCHE and E represents PE [114]. The resulting polymers showed good processibility

and unexpectedly high toughness. Furthermore, a group of heptablock terpolymers CEC-P-CEC,

in which P represents poly(ethylene-alt-propylene) in the center, was prepared. These materials

demonstrated tensile strength around 30 MPa and strain at break higher than 500% [115].

3.1.2 Polymers prepared by cationic polymerization

Another approach to improve the UV and oxidative stability of SBCs is to incorporate

polyisobutylene (PiB) as a saturated non-crystalline elastic block to replace the polydiene. A


25

renaissance in “living” cationic polymerization advanced research towards PiB based TPEs with

better UV/oxidation resistance and higher UST [28]. Triblock copolymer PS-PiB-PS prepared by

sequential living cationic polymerization using a difunctional initiator exhibits an ultimate tensile

stress of 26 MPa, which is competitive with commercial Kraton SIS TPEs [116, 117]. The

molecular weight of PiB should be targeted between 40,000 to 160,000 g/mol in order to achieve

optimal stress-strain performance. Phase separation is observed between PS and PiB when the Mn

of PS is above 5,000 g/mol. Other triblock copolymers with PiB as the elastic middle block and

glassy end blocks of polystyrene derivatives, such as poly(tert-butyl styrene) [118] and poly(α-

methyl styrene) [119-121] were synthesized by living cationic polymerization. PMMA-PiB-

PMMA triblock copolymer was prepared by a cationic/anionic mechanism switching process

[122]. All these triblock copolymers demonstrated similar stress-strain behavior as compared to

TPEs prepared by anionic polymerization with polydienes as the rubbery matrix. High Tg

polymers, polyindene (PID, Tg ~ 209 °C) [123] and its derivative polyacenaphthylene (PACP, Tg

~ 250 °C) [124], have been employed successfully with PiB as the elastic block. Poly(p-

chlorostyrene) (PpCS, Tg ~ 129 °C) [125], a polar polymer with flame resistance, was also

successfully incorporated as the hard segment of PiB based TPEs. Alloocimene is produced by

thermal isomerization of α-pinene, a cyclic terpene from sustainable resources, and is subject to

cationic polymerization. Puskas synthesized a group of triblock copolymers with PiB as the middle

elastic block and polyalloocimene (Tg ~ 71oC) as hard block. With 23.7 wt% of the glassy block,

the polymer displays 14.5 MPa ultimate strength and 743% elongation at break [126-128].

Alkyl vinyl ethers are another unique class of monomers that are cationically polymerizable.

Well-defined poly(vinyl alcohol) (PVA) was produced by the hydrolysis of cationically

synthesized poly(tert-butyl vinyl ether) (PtBVE, Tg ~ 74°C) [129]. As a widely used water-soluble
26

polymer in coating and textiles, semicrystalline PVA has a Tg of 80 °C and Tm of 208 °C [130],

which makes it a suitable candidate for TPEs with oil resistant properties. Attempts at preparing

PVA-PiB-PVA triblock copolymer by direct hydrolysis of PtBVE-PiB-PtBVE were unsuccessful

due to solubility issues [131]. Cyclohexyl vinyl ether (CHVE) could be statistically copolymerized

with tBVE by living cationic polymerization initiated from the difunctional polyisobutylene living

cation. Selective hydrolysis of PtBVE in triblock copolymers yielded P(CHVE-stat-VA)-PiB-

P(CHVE-stat-VA) triblock copolymers containing a statistical copolymer of PCHVE and PVA as

the rigid phase. Triblocks with 25 -75 mol% of VA in the hard block were still soluble in THF.

PVA containing triblock copolymers exhibited an increase in mechanical strength from 15.9 to

22.6 MPA with an increase of VA content [131].

Similar to poly(alkyl acrylate)s, the Tg of poly(alkyl vinyl ether)s are tunable by adjusting the

length of alkyl substituents [132]. With poly(n-butyl vinyl ether) (PnBVE, Tg ~ -55 °C) as the

elastic block, ABA-type triblock copolymer with poly(tricyclodecane vinyl ether) (PTCDVE, Tg

~ 77 °C) [132], and poly(2-adamantyl vinyl ether) (PADVE, Tg ~ 150 °C) [133] as glassy blocks

were synthesized by living cationic polymerization. All these vinyl ether TPEs showed relatively

low strain at break (<355%) and low ultimate stress (<5.55 MPa), which might be attributed to the

high Me of PnBVE. Hashimoto prepared a unique group of all-vinyl ether triblock copolymers by

cationic polymerization with PADVE as the rigid block and poly(vinyl ether) with different polar

groups as elastic blocks. Although the mechanical performance of all-vinyl ether TPEs is modest,

these polymers demonstrated excellent CO2 separation ability, which is close to Robeson’s upper

bound (Figure 9), by incorporating inner blocks with various polar functional groups, such as

ester, hydroxyl, and oxyethylene units. [132, 134, 135]. The permeability of CO2 drastically

increased to 300 barrier as the weight percentage of oxyethylene-based soft segment increasing.
27

While physical crosslinks of the glassy segments provided the mechanical strength of the material

in the triblock copolymer type CO2 separation membrane, the rubbery segments with higher free

volumes formed microdomains for gas permeation [136].

Figure 9. Plots of PCO2/PN2 vs. PCO2 for all-vinyl ether triblock copolymers with different

number of oxylethylene units in the middle repeat unit together with related polymers for

comparison [134]. Copyright 2017, Reproduced with permission from Springer-Verlag Berlin

Heidelberg. The solid line shows Robeson’s upper bound (2008) [135].

3.1.3 Polymers prepared by ring-opening polymerization (ROP):

The unique advantage of adopting ROP to prepare TPEs is that various monomers suitable for

ring opening polymerization could potentially be derived from sustainable resources, which has

been a long-standing goal of preparing polymers from non-petroleum resources [137]. Tang [138]

and Hillmyer [139] have summarized recent advance in preparing sustainable polymers from

renewable resources. Monomers including racemic D,L-lactide, 3-hydroxybutyrate (HA), menthide

(MD), 6-methyl-ε-caprolactone (MCL), caprolactone (CL), β-methyl-δ-valerolactone (MCL), ε-


28

decalactone (DL) could all be polymerized by ROP [140]. In some pioneering research to prepare

poly(lactide) (PLA) containing triblock type TPEs, ionically prepared α,ω-dihydroxyl PI [141],

PiB [142], polydimethylsiloxane (PDMS) [143], polyethylene oxide (PEO) [144] and PEP[145]

were employed as the macroinitiators for the polymerization of poly(lactide). However, the elastic

blocks of these copolymers are not derived from sustainable resources or biodegradable. The

section here focuses on ABA triblock type TPE materials derived from renewable resources.

The most commonly employed rigid block in ROP routes to TPEs is poly(lactide), an amorphous

biodegradable polymer with a Tg around 60 oC. The stereocomplex crystals formed by equimolar

racemic blending poly(D-lactide) (PDLA, Tm ~ 175 °C) with semicrystalline poly(L-lactide)

(PLLA, Tm ~ 175°C) can further increase the melting temperature [137]. Additionally, the

stereocomplex improved the tensile strength, modulus, elongation at break, and heat and enzymatic

degradation resistance of PLA as compared to PLLA or PDLA [146]. These improvements were

attributed to tighter molecular packing and smaller molecular mobility despite the similar chain

conformation in PLA stereocomplex (Figure 10) [147]. Hillmyer took full advantage of

stereochemistry and prepared a series of biodegradable TPEs with polymenthide (PMD, Tg ~ -

25oC) as the elastic block and either PLA, PLLA or PDLA as the hard block [147-149]. The

presence of semicrystalline PLLA or PDLA in the hard blocks enhanced the physical crosslinking,

resulting in samples with performance similar to styrenic TPEs. Blends of the enantiomeric

triblock copolymers lead to extensively increased stiffness, strength, and elongation at break [149].

PDLA-PMD-PDLA triblock copolymers could also serve as triblock nucleating agents to blend

with commercially relevant PLLA.


29

Figure 10. Schematic representation of the nucleation of PLLA using PDLA–PM–PDLA

triblock copolymers [147]. Copyright 2009, Reproduced with permission from John Wiley and

Sons Inc.

For elastic blocks, Xiong developed an economically viable strategy to prepare β-methyl-δ-

valerolactone (βMδVL, Tg ~ -51 °C) through an artificial biosynthetic approach [150]. Excellent

tensile properties were obtained by utilizing PβMδVL as the elastic block with PLA as the hard

segment. With 32 vol% of PLA, PLA-PβMδVL-PLA triblock copolymer displayed 1725 % strain

at break with ultimate stress of 25 MPa, which is competitive with commercial Kraton copolymer

based on petroleum resources. However, the six-member ring structure in βMδVL led to a low

polymerization rate and low ceiling temperature for polymerization. Incomplete reaction at

equilibrium conditions and the resultant requirement of removal of residual monomers has limited

further application of βMδVL based polymers [151]. Hillymer demonstrated that for lactones with

the same ring strain (the same number of atoms in the ring), the position and length of substitution

groups on lactone rings played a significant role on the polymerization kinetics [152].

Another commonly utilized elastic block for ROP prepared TPEs is poly(ε-caprolactone) (PCL),

a biodegradable semi-crystalline material with a low Tg (Tg ~ -60 °C, Tm ~ 60 °C). Copolymerizing

ε-caprolactone with ε-decalactone or lactide creates amorphous soft segments suitable for TPE
30

applications [153, 154]. With 30 vol% of poly(6-methyl-ε-caprolactone) (PMCL, Tg ~ -45 °C) as

the elastic block, 1880 % strain at break was achieved with 10.2 MPa ultimate stress [155].

Statistical copolymer constructed from ε-caprolactone and D-lactide showed a Tg of -30 °C. With

20 wt% of isotactic poly(L-lactide) (PLLA) as the glassy block, triblock copolymers showed

exceptionally high strain at break of 2800 % and fair ultimate stress of 17 MPa [152]. By varying

volume fraction of the hard segment PLA from 17 % to 32 %, TPEs based on a random copolymer

of ε-caprolactone/ε-decalactone as the soft segment demonstrated tunable ultimate stress from 9.9

MPa to 18 MPa with strain at break from 1200 % to 2100 %. With close structural similarity,

poly(γ-methyl-ε-caprolactone) (PγMCL, Tg -60 oC) has also been employed as the soft segments

in TPEs. For polymers containing 17 vol% of PγMCL, excellent mechanical properties (30 MPa

ultimate stress with 988% strain at break) were observed. It is worth noting that the Me of PγMCL

was 2.9 kg/mol, which is comparable to PB (Me = 1.7 kg/mol) and lower than PI (Me = 6.1 kg/mol)

[151].

3.1.4. Polymers prepared by controlled radical polymerization techniques:

Atomic transfer radical polymerization (ATRP), reversible addition-fragmentation chain

transfer polymerization (RAFT) and nitroxide-mediated radical polymerization (NMP) are three

major controlled radical polymerization (CRP) techniques [156]. These techniques have opened

various opportunities to prepare a wider range of functionalized polymers with predictable

molecular weight, narrow molecular weight distribution and complex macromolecular

architectures [157, 158]. As compared with ionic polymerization, CRPs have higher tolerance to

oxygen and moisture, and fewer limitations with regard to the choice of monomers. Controlled

polymerizations have been achieved for monomers such as acrylonitrile [159], acrylamide [160],

and vinyl amide [161], monomers which are not polymerized in a controlled fashion by ionic
31

mechanisms. For example, poly(α-methylene-γ-butyrolactone) (PMBL) is a rigid thermoplastic

with a Tg of around 195 °C [162]. The renewable monomer MBL can be derived from tulipalin A

and subjected to radical polymerization [163]. A group of PMBL-PM-PMBL triblock type TPEs

with poly(menthide) as the elastic block were prepared for high-temperature applications [164].

Ultimate tensile stresses ranging from 3.9 MPa to 13 MPa and strains at break ranging from 730

% to 1800 % were achieved based on different volume fractions of PMBL. Derivatives of itaconic

acid such as N-phenylitaconimide and N-(p-tolyl)itaconimide can be produced from biomass and

are suitable for RAFT polymerization [165]. Triblock copolymer with 27 mol% of poly(N-

phenylitaconimide) as the hard block and 73 mol% of poly (n-butyl itaconate) as soft block

displayed potential application for high temperature thermoplastic elastomers with a Tg up to 241
oC.

One issue associated with classical CRP techniques is the limited capacity to maintain the chain

end integrity at 100% conversion. Terminating the reaction before quantitative monomer

consumption is required in order to prepare well-defined block copolymers. To alleviate this issue,

difunctional initiators or chain transfer agents (CTAs) have been commonly employed to

simultaneously propagate hard blocks in the chain end of an ABA triblock copolymer. In ATRP,

1,2-bis(bromoisobutyryloxy) ethane (bBr-iBE) [166] and diethyl meso-2,5-dibromoadipate

(DEMDBA) [84] were utilized to synthesis PMMA-b-PnBA-b-PMMA triblock copolymers. In

NMP, a difunctional alkoxyamine initiator with N-tertbutyl-1-diethylphosphono-2,2-

dimethylpropyl nitroxide (SG1) was utilized to synthesize PS-PnBA-PS triblock copolymers from

the middle block [167, 168]. For RAFT polymerization, different types of difunctional CTAs were

used to initiate the polymerization either from a middle soft block (CTA-B-CTA type approach)

by using CTAs like 1,4-bis(thiobenzoylthiomethyl)benzene (BTBTMB) [169] or from the hard


32

block (A-CTA-A type approach) by using CTAs like dibenzyl trithiocarbonate (DBTTC) [170].

Arkema developed a monofunctional NMP initiator approach based on BlocBuilder® alkoxyamine

(SG-1, Figure 11.a) to prepare Nanostrength®, a PMMA-PnBA-PMMA triblock copolymer [33].

This initiator demonstrated advantages including excellent control over the polymerization of

acrylates and methacrylates, fast reaction, and no metal residue in the product. A modified

difunctional initiator (DIAMA) from SG-1 (Figure 8.b) to build symmetric triblock copolymers

by NMP was also reported (Figure 11.b) [34].

Figure 11. Initiators for the synthesis of PMMA-b-PnBA-b-PMMA triblock copolymers via

NMP. (a) SG-1, (b) SG1- based difunctional alkoxyamine (DIAMA) [168].

Compared with conventional free radical and ionic polymerization, CRP generally requires

lower active center concentrations to minimize undesired chain transfer and termination. However,

very low radical concentrations lead to longer reaction times, which increases the production cycle

in industrial scale processes. Radical segregation effects introduced by (mini)emulsion

polymerizations in a heterogeneous system, on the other hand, reduced the reaction time and

suppressed radical termination [171, 172]. PS-PnBA-PS triblock copolymers with different

molecular weights and compositions were prepared in a shorter reaction time by combining

emulsion polymerization with RAFT polymerization [98]. By varying weight percentage of PS

from 20.2 % to 71.5 %, ultimate tensile strength ranged from 3.0 MPa to 12.5 MPa and strain at

break was in the range from 90 % to 1300 %. PS-PnBA-PS block copolymer were also prepared
33

with a poly[styrene-alt-(maleic anhydride)] (PSM) macro-chain transfer agent in miniemulsion

polymerization [172]. SBS triblock copolymers were prepared by sequential addition with 1-

phenylethyl phenyl dithioacetate as CTA. Surprisingly, the elastic PB block contained 12% vinyl-

1,2 structure, 50% trans-1,4 structure, and 38% cis-1,4 structure, which was similar to PB prepared

by anionic polymerization in hydrocarbon solvent [173]. Similarly, another TPE based on PS and

poly(lauryl acrylate) was prepared by a RAFT polymerization process [169]. Ultimate stress was

lower than 1 MPa and strain at break was lower than 280 %.

Despite success in suppressing the undesired chain transfer and termination and reducing the

producing time, the mechanical properties of TPEs prepared by CRP still cannot compete with

those prepared by living anionic polymerization. Tong compared the mechanical properties of two

PMMA-PnBA-PMMA triblock copolymers prepared by ATRP and living anionic polymerization.

These two polymers had the same molecular characteristics except for the PDI of the sample

prepared by ATRP (PDI: 1.15) was 10% higher than that prepared by LAP (PDI: 1.05). In the

dynamic mechanical analysis, the plateau of storage modulus of the LAP sample was a decade

higher than for the ATRP sample. No sharp drop of the storage modulus in the high temperature

region was observed for the product made by ATRP, indicating the existence of more diblock

species (with a higher hard block ratio) and weaker phase transition as compared to the product

made by LAP. A broad transition peak in the temperature range from 0 to 80 oC was also observed

in the tan δ curve. In tensile tests, the LAP sample had a higher initial modulus, and higher stress

and strain at break than the ATRP sample (as shown in Figure 12) [85]. The difference in the

mechanical performance was attributed to the relatively broad molecular weight distribution of the

PMMA and the presence of unreacted PMMA-PnBA diblock copolymer in the ATRP sample.

Broader molecular weight distribution appears to reduce the strength of phase separation, with the
34

unavoidable diblock copolymer in the triblock copolymers acting as a plasticizer [66, 74].

Differences in mechanical performance were also observed in PMMA-PnBA-PMMA triblock

copolymer supplied by two commercial sources: Arkema’s Nanostrength® was synthesized by

NMP with tensile strength lower than 1 MPa [33]. Kuraray’s KurarayTM was synthesized by living

anionic polymerization with tensile strength around 8 MPa [34].

Figure 12. Mechanical properties of PMMA-b-PBA-b-PMMA triblock copolymers synthesized

by living anionic polymerization (LAP) and atom transfer radical polymerization (ATRP): (a)

Storage modulus as a function of temperature; (b) stress-strain curves [85]. Copyright 2000,

Reproduced with permission from the American Chemical Society.

Compared with living anionic polymerization, CRP has limited capacity to achieve the standard

of “living” polymerization, especially for classical monomers such as isoprene or butadiene.

However, the high tolerance of CRP to various functionalities have enabled chemists to access

novel materials. By taking advantage of both CRP and supramolecular chemistry, smart TPEs with

novel properties, such as self-healing and shape memory, have been prepared. Details on these

developments are presented in section 5: TPEs based on supramolecular chemistry.

3.1.5. Polymers prepared by living coordination polymerization techniques:


35

Polyolefins represent almost two-thirds of global consumption of plastic and elastomer

products. Ethylene and propylene are the two main feedstocks for polyolefins due to their low cost

and wide availability. Linear polyethylenes (PE), as well as isotactic and syndiotactic

polypropylene (i-PP and s-PP), are semi-crystalline polymers that could be employed as the hard

segments for TPE applications. Atactic polypropylene (a-PP) and poly(ethylene-co-propylene)

(PEP) are amorphous soft polymers that may be suitable for elastic segment [174]. Combining

these two types of polymers has resulted in olefin-based block copolymers that are potential low-

cost alternatives to hydrogenated SBCs TPEs.

Natta envisioned a possible TPE structure based on stereoblock homopolymers with i-PP as the

hard segment and a-PP as the soft segment in 1959 [175] and prepared the corresponding material

by fractionation. Collette first reported PP stereoblock homopolymers synthesized with alumina

supported tetra-alkyl group IVB catalyst [176, 177]. The elasticity of PP was attributed to the co-

crystallization between PP with different stereoregularities, where the high molecular weight a-PP

formed the soft matric in the polymer network, and the stereoregular i-PP formed the

semicrystalline physical crosslinking junctions. Chien reported an ansa-metallocene catalyst that

polymerized propylene into crystalline-amorphous block TPEs [178]. The material demonstrated

typical TPE behavior with 12.1 MPa strength and 1260% strain at break. By using “dual-side”

zirconocene catalysts, Dietrich prepared polypropylenes having various i-PP and a-PP contents. A

continuous change of the mechanical properties from elastomer to thermoplastic was achieved

[179]. Waymouth reported an atactic-isotactic stereoblock polypropylene made by using

oscillating stereocontrol. However, the relatively high Tg of atactic PP (Tg ~ 0 oC) limits the

applications [180]. Rieger demonstrated the synthesis of PP based TPEs with i-PP and a-PP block

structures by using ethylene-bridged fluorenylindenyl (EBFI) ansa-metallocene complexes as the


36

single catalyst [181]. Eisen evaluated the capacity for modulating the stereo structure of PP by

rapidly changing the monomer feeding pressure [182].

Compared with conventional Ziegler-Natta catalysts, the use of “living” Ziegler-Natta catalysts

provides better control of the macromolecular architecture, stereo-regularity and the composition

[183, 184]. Sita used a specifically designed zirconium/borate complex for programmable stereo-

modulated “living” Ziegler-Natta polymerization [185]. Simply by adjusting the zirconium/borate

ratio and disrupting propylene feeding during the polymerization, diblock, triblock and tetrablock

copolymers consisting of a-PP and i-PP were prepared with 40 wt% of i-PP. Tetrablock PP

demonstrated 1227% strain at break with a stress around 20 MPa. Coates prepared various olefin

TPEs with metal complexes catalysts that were different from Sita [186]. With titanium catalysts,

sPP-b-PEP-b-sPP (syn-TB) triblock copolymers were synthesized and displayed a cylindrical

morphology with 25 wt% s-PP. By using nickel catalysts, iPP-b-aPP-b-iPP (iso-TB) triblock

copolymers and iPP-b-aPP-b-iPP-b-aPP-b-iPP (iso-PB) pentablock copolymers were synthesized

with propylene monomer. Despite a disordered morphology, iso-PB exhibited exceptional

mechanical behavior with a strain at break of 2400% and tensile strength of 250 MPa. Note that

all the mechanical strengths reported in this work were true stress instead of the more commonly

used engineering stress. The advantages of using semicrystalline polyolefins as TPEs are the

superior chemical resistance and lower cost of the monomers. Additionally, for low molecular

weight samples, the large enthalpy of crystallization can overcome the small favorable free energy

of phase mixing. Thus, designing TPEs with lower molecular weight is possible, which might

bring benefits during melt processing.

Other than using PE or P-iPP as the rigid block, Sita reported another triblock copolymer type

TPEs based on poly(1,3-methylenecyclohexane) (PMCH, Tg ~ 92 oC, Tm ~ 209 oC) [187] as the


37

hard domains [188]. Triblock copolymers of PMCH-b-a-PP-b-PMCH were prepared by LCP with

hafnium catalysts (Figure 13). The rigid block PMCH was formed by ring closing of 1,6-

heptadiene, and the resulting polymer PMCH was micro-phase separated from a-PP. As the wt%

of PMCH increased from 9% to 23%, the ultimate stress increased from 8.9 MPa to 20.3 MPa, and

the elongation at break decreased from 2773% to 1390%. The triblock copolymer displayed

almost perfect elastic recovery.

Figure 13: Stress-strain behavior and corresponding morphology of triblock TPEs based on

PMCH and a-PP [188]. Copyright 2015, Reproduced with permission from the American

Chemical Society.

Another elegant innovation with which to prepare polyethylene and poly(α-olefin) copolymer

based TPEs is chain straightening/walking polymerization by using late transition metal complexes

with the α-diimine complex as the catalyst [189]. Rather than adding incoming monomers at the

chain end, the chain walking mechanism assembles the monomers on the polymer backbone. In
38

this case, the polymerization of ethylene or propylene generates branched amorphous segments,

whereas the polymerization of higher α-olefins generates linear semicrystalline segments [190-

192]. Leone and Ricci synthesized triblock copolymer with poly(1-dodecene) as the rigid block

and branched polyethylene as the elastic block. The triblock copolymer demonstrated tensile

strength of 18 MPa with ultimate stress of 1000%. However, a permanent deformation around 175%

was observed in the 1st hysteresis experiment. Coates employed a Ni(II) catalyst with a different

ligand and prepared a similar triblock copolymer with improved elastic recovery [192]. Chen

demonstrated the ability to use only ethylene as the monomer feedstock to prepare TPEs [191].

3.2 ABC triblock terpolymers type TPEs

In traditional ABA triblock copolymers type TPEs, the thermodynamic incompatibility between

A and B blocks leads to the microphase separation of soft matrix and hard physically crosslinking

junctions. Spherical or cylindrical morphologies are usually formed with a low ratio of the hard

block (block A). Polymer chains of the B block can either form loops placing A blocks within the

same domain (Figure 14.a), or form bridges tethering A blocks between two different hard

domains. (Figure 14.b) [193]. Only the latter arrangement contributes significantly to elastic

recovery. Different methods, such as dielectric relaxations [193] and dynamic viscoelastic

measurements [194, 195], have been used to determine the bridge fractions (ϕbridge), for which the

theoretical ϕbridge for lamellar, cylindrical and spherical forming ABA triblock copolymer are 40%

- 45%, 60% - 65%, and 75% - 80%, respectively [193, 196, 197]. With a similar level of

entanglement in the elastic domain, higher ϕbridge has been demonstrated to be beneficial to better

mechanical performance, including modulus, stress at break and elasticity. Suppressing the

formation of loops and enhancing the formation of bridges in elastic blocks have the potential to

improve the mechanical properties of TPE materials. This leads to the design of ABC triblock
39

terpolymer type TPEs, in which all three blocks are immiscible with each other and form distinct

domains thus promoting bridge formation (Figure 14.c).

Figure 14. Schematic illustration of polymer chain conformations in ABA triblock copolymer:

(a) Loop-B; (b) Bridge-B [193]. Copyright 1995, Reproduced with permission from the American

Chemical Society. (c) Schematic illustration of polymer chain conformations in ABC triblock

copolymer.

Various attempts have been made to create such ABC triblock terpolymers type TPEs.

Brinkmann-Rengel synthesized both PS-b-PB-b-PMMA (SBM) and PS-b-PB-b-PS (SBS) triblock

copolymers with similar block ratios and molecular weights by “living” anionic polymerization

[198]. Since the thermodynamic incompatibility between PS and PMMA was weak (low χ),

suppression of loop formation only appeared at high molecular weight (large N). The segregation

of all blocks was observed for high molecular weight PS and PMMA in both SBM and

hydrogenated SEBM [199-201]. Deformation occurred in the M phase for lamellar forming SBM

terpolymers but in the S phase for cylinder forming terpolymers [202]. Additionally, the molecular

weight and the microstructure of PB blocks also influenced the mechanical properties of the

terpolymers [203]. Compared with PMMA, poly(2-vinyl pyridine) (P2VP) has a larger χ with

respect to both PS and PI. Polystyrene-b-polyisoprene-b-poly (2-vinyl pyridine) (SIP) exhibited a

higher modulus than SIS with similar composition and molecular weight [194, 195].
40

In diblock copolymers, incorporating semicrystalline blocks could be used to induce strong

phase separation even at relatively low molecular weights [204-206]. By introducing one

semicrystalline end block in an ABC triblock terpolymer, several different TPE materials have

been prepared [207]. Stadler and co-workers synthesized polystyrene-b-polybutadiene-b-

poly(caprolactone) (PS-b-PB-b-PCL) triblock copolymers through sequential anionic

polymerization in benzene [208]. The mechanical properties and morphology of the triblock

terpolymers was also strongly affected by the specimen preparation methods [209, 210].

Abetz and co-workers evaluated the influence of crystallization on the mechanical properties of

TPEs based on hydrogenated SIS and polystyrene-b-polyisoprene-b-polybutadiene (SIB) triblock

polymers. The hydrogenation converted PI into poly(ethylene-stat-propylene) (PEP, Tg: ~ -56 oC)

and PB into polyethylene (PE, Tm: ~ 88 oC), resulting in SEPS and SEPE triblock copolymers.

PEP is an amorphous polymer that serves as the elastic segment; linear PE is a semicrystalline

polymer which serves as the rigid end block [211]. In contrast to the cylindrical morphology

formed in SIS, SEPE exhibited a morphology consisting of PS cylinders and PE crystallites within

a matrix of the PEP block [212]. At lower strain (less than 300%), SEPE showed better elasticity;

at higher strains, SEPS showed better elastic recovery behavior [211, 212]. The morphology

dependence of deformation provided a rational interpretation of this phenomenon: at lower strain,

the semi-crystalline structure of PE and the increased bridge fraction contributed to the better

elasticity of SEPE. However, at higher strain, the crystalline PE was easier to be irreversibly pulled

out from the crystalline domains than the amorphous PS, leading to the reduced elastic recovery

of SEPE [213].

3.3 Multiblock polymer type TPEs


41

Different from ABC triblock terpolymers where various new morphologies were observed due

to the thermodynamic incompatibility, between end blocks as well as midblocks, present in these

systems, linear alternating (AB)n multiblock copolymers could also lead to profound impact on the

physical properties without significantly altering the phase behavior [214]. Comparing the polymer

architectures of AB diblock, ABA triblock, and ABABA pentablock copolymers where all

polymers exhibit lamellar morphologies, the knotted looping structures presented in the center A

block of ABABA pentablock copolymers “stitched” the domains (Figure 15) and enhanced the

mechanical resistance to deformation and cracking [215]. Blending only 10-15 wt% of CECEC

pentablock copolymer into CEC triblock copolymers resulted in a material that exhibited similar

morphological and mechanical characteristics to the pure CECEC pentablock copolymers [215].

Figure 15. Schematic illustration polymer chain conformation of AB diblock, ABA triblock,

and ABABA pentablock copolymers with lamellae morphology [215]. Copyright 2003,

Reproduced with permission from the American Chemical Society.

Bates investigated the phase behavior and viscoelastic properties of various symmetric (AB)n

multiblock copolymers synthesized by a combination of living polymerizations and coupling

reactions. These polymers consisted of PS, PCHE, PE or PLA as the rigid A block, and PB or PEP

as the elastic B block [115, 216-221]. For (PE-PEP)n multiblock copolymers, an intriguing
42

transition from strain softening to strain hardening was observed between n = 8 and n = 10 [217].

A similar transition was also observed in semicrystalline PLA containing multiblock copolymers

[220]. The author attributed this transition to the reduction of nucleation rate and polymer chain

mobility. As the number of repeating units (n) increases, the mobility of polymer chains is reduced

since the interior blocks are locked between two interfaces, leading to a reduced rate of

crystallization. The multiblock architecture could dramatically enhance the ultimate stress and

strain with minor impact on the yield stress. Even though these results were based on symmetric

compositions such that most of these materials were considered as thermoplastic, the strategy

certainly provides insights for investigating the structure-property relationships and improving the

mechanical performance of segmented TPEs such as polyurethanes (TPU), copolyesters (COPEs),

and copolyamides (COPAs).

Most of the segmented multiblock copolymers are prepared by step-growth polymerization

through α,ω-hydroxyl or isocyanate functionalized polymers or oligomers. Recently, several

groups adopted thiol click reaction to prepare multiblock copolymers for thermoplastic and/or

elastomer applications [222, 223]. Compared with moisture-sensitive urethane chemistry, click

chemistry has higher tolerance to functional groups, water, and oxygen. Becker and Dove prepared

a group of elastomers using organocatalyzed stereo-specific addition of thiols and activated

alkynes. The ability to tune the percentage of cis-double bonds through solvent polarity was the

key to preparing elastomers with a wide range of mechanical properties (Figure 16) [224]. These

workers used 1,6-hexanedithol and ethylene glycol dimethacrylate to prepare the soft block and

then extended the chain with N,N`-methylenebis(acrylamide) to prepare the multiblock TPEs in

two steps. The incorporation of acrylamide moieties provided hydrogen bonds that physically

crosslinked the network [225]. Long took advantage of the inherent chemoselectivity of thiol-
43

Michael addition and prepared a poly(β-thioester) containing TPEs in one-pot with bis-oxamide

as the donor for four hydrogen bonds [226]. Thiol-norbornene and thiol-maleimide reactions were

also employed to prepare thermoplastics [227].

Figure 16. Organocatalyzed stereo-specific addition of thiols and activated alkynes [224].

DMF/CHCl3 (100) refers to 100% of DMF as the solvent. Copyright 2016, Reproduced with

permission from John Wiley and Sons.

Chain shuttling polymerization was designed by Dow Chemical Company to prepare multiblock

olefin copolymers with alternating semicrystalline and amorphous segments. The polymerization

involved a chain shuttling agent (CSA) that passing the growing polymer chains between two

catalysts that have stereo- or monomer selectivity [228-230]. When the polymerization was

conducted without CSA, a blend of semicrystalline PE and amorphous PE was obtained. With the

presence of CSA, multiblock PE with wide range access of mechanical properties was prepared

on an industrial scale [228]. Zinck prepared multiblock TPEs by chain shuttling polymerization

with isoprene and styrene as the monomer pair [231].

Other than above-mentioned strategies, many novel polymerization/coupling methods have

been developed to prepare multiblock copolymers. Ling combined cationic and anionic ring-

opening polymerization and prepared polycaprolactone-polytetrahydrofuran multiblock


44

copolymers in a “Janus Polymerization” fashion. The resulting material displayed around 1300%

strain at break with ultimate stress above 10 MPa [232]. While the conversion of caprolactone

could achieve over 80%, the maximum conversion of tetrahydrofuran was 36.8%. The same group

prepared multiblock copolymers with entirely degradable polyester or polypeptide backbone by

using a L-lysine diisocyanate coupling agent [233]. Bazan took advantage of cross acrylic diene

metathesis of α,ω-divinyl-terminated polyolefins and prepared multiblock semicrystalline/

amorphous multiblock copolymers using a one-pot protocol [234]. Lee prepared polystyrene-

polyolefin multiblock copolymers by sequential coordination and anionic polymerization [235].

4. Branched copolymers for TPEs

Compared with linear block copolymers, branched polymers have unique features including

smaller hydrodynamic radii, and potentially lower melt viscosity and lower solution viscosities

[236]. For application as TPEs, branched copolymers offer an additional parameter,

macromolecular architecture, that can be used to manipulate properties and potentially decouple

the morphological dependence on composition that is observed for linear TPEs. Here, we focused

on star, graft and hyper-branched macromolecular architectures.

4.1 Star branched copolymers for TPEs:

Star-branched polymers are polymers with more than two “arms” radiating from the same core.

If these arms have different chemical compositions or molecular weights, the star polymer is

named as a miktoarm (mixed-arm) star polymer (Figure 17) [237, 238]. There are two common

methodologies to build the star architecture: (1) “Arm-first” where polymer arms are synthesized

first and linked to a core decorated with appropriate reactive sites. (2) “Core first,” where polymer

arms are grown from a multi-functional initiator. The star-shaped TPEs are usually composed of
45

soft domains as the inner block and hard domains as the outer block. PS/PI and PS/PB star-block

copolymer TPEs are commercially available under the trade name Kraton®.

Figure 17. Illustrations of symmetric and asymmetric star copolymers.

4.1.1 Polymers prepared by anionic polymerization

Synthesized by living anionic polymerization, star copolymers composed of PS, and PB or PI

have been extensively studied due to their well-defined structures. When more than two PS-b-PI

diblock copolymers are connected at the same core through the PI end with proper composition,

the resulting (PS-b-PI)x star-branched polymers display mechanical properties similar to SIS linear

triblock TPEs. By using an arm-first approach and divinylbenzene linking strategy, Fetters

prepared polystyrene-polydiene star-block copolymers with up to 29 arms [239]. These star

polymers had superior tensile stress and modulus as compared to linear triblock copolymers of

similar composition. The enhancement of tensile strength saturated when the number of arms was

higher than six. Morphological analysis indicated that with the same molecular weight, multi-arm

star polymers had smaller PS domains size as compared with linear polymers [240]. Thus, star

polymers had more condensed physical crosslinks per unit volume, leading to higher tensile

strength. Additionally, the core in star polymers acted as permanent crosslinks due to the covalent

chemical linkage and further enhanced the tensile strength. In addition to the higher tensile strength

of star polymers, the intrinsic viscosity of star polymers was lower than their linear analogs.
46

Spencer and Matsen used self-consistent field theory to calculate the bridging statistic for star

copolymers. For (AB)m star block copolymers with m AB diblock copolymers attached at the B

end, virtually all B blocks formed bridge when m is higher than 10 [241].

Dair et al. reported the characterization of PS-b-PI star copolymers exhibiting a double gyroid

morphology [242]. Compared to materials with spherical, cylindrical, and lamellar morphologies,

the 3-dimensionally interpenetrated periodic structure prevents the sliding of the junctions while

stretching. Thus, the resulting polymer exhibited yield stresses considerably higher than for

classical SBCs. However, the strain was only around 350% to 400%, which was much lower than

for conventional SIS TPEs.

Confirmed by both experiments [243] and theory [244], the morphological dependence of block

copolymers can be decoupled from the chemical composition by varying macromolecular

architecture. Progress in self-consistent field theory [245] facilitated the advanced design of TPEs

based on nonlinear architectures such as miktoarm star copolymer with superior mechanical

properties [246]. For linear SIS triblock copolymer, over 36 vol% of PS component leads to a

lamellar morphology which is generally unfavorable for TPE applications [247]. For A(BA’)4

miktoarm star polymer with one A block and four BA’ blocks emanating from the same core,

Fredrickson predicted a stable cylindrical morphology [246]: A phase hexagonally dispersed in B

matrix with the volume fraction of A polymer up to 70 %. As for experiments, miktoarm star

copolymer S(IS’)3 with 50 vol% of PS achieved stable cylindrical morphology (Figure 18) [248].

The high-volume fraction of PS in this type of asymmetric miktoarm PS-b-PI copolymers enabled

a new form of TPE with a higher modulus, strength, toughness and recoverable elasticity [248,

249]. The SIS’ with 50 vol% of PS yield at low elongation, indicating its thermoplastic nature. By

blending with PS homopolymers, a new stiffer TPE (modulus was 99.2 MPa) with aperiodic
47

“bricks and mortar” mesophase morphology was achieved with up to 82 wt% of PS [250]. Using

similar miktoarm star copolymer by blending with PS, Shi created a lamellar morphology with up

to 97 wt% of PS [251].

Figure 18: S(IS’)3 miktoarm star copolymer type TPEs [248]. Copyright 2014, Reproduced

with permission from the American Chemical Society.

For the “core-first” strategy: developing multifunctional anionic initiators received limited

success mainly because of the poor solubility of such initiators in hydrocarbon solvents [252].

Baskaran prepared a trifunctional organolithium initiator. The novel initiator was used to prepare

PI-PS three-arm star polymers by sequential anionic polymerization in hydrocarbon solvent

without polar additives [253]. However, since the star-block copolymer was connected at the PS

chain ends, mechanical properties were not investigated.

4.1.2 Polymers prepared by cationic polymerization

While star copolymer TPEs were rarely prepared using multifunctional anionic initiator by the

“core-first” strategy, many star copolymer TPEs have been prepared using multifunctional cationic

initiators. (PpCS-PIB)8 octa-arm star copolymers were prepared through a calix[8]arene core with

eight initiation sites[254]. (PMMA-PIB)3 three-arm star copolymers were prepared by a

trifunctional cationic initiator followed by ATRP of MMA [255]. For the “arm-first” strategy: at

the end of “living” cationic polymerization, vinyl functionality was introduced by reacting the
48

“living” cation of PS-PIB+ and PID-PIB+ with allyltrimethylsilane. The vinyl end functionality

was further reacted with Si-H on cyclosiloxane by Pt-catalyzed hydrosilylation and produced star-

block copolymers of (PS-PIB)n and (PID-PIB)n based on different numbers of Si-H on

cyclosiloxane [256-258].

Similar to the arm first divinylbenzene linking strategy for anionic polymerization, 1,4-

cyclohexane dimethanol divinyl ether was applied as the linking agent for arm first cationic

polymerization in order to prepare star copolymers with poly(2-adamantyl vinyl ether) as the hard

segment and poly(n-butyl vinyl ether) as the elastic segment. Both linear and star types were

synthesized through living cationic polymerization. Both the tensile strength and elongation values

for star type copolymers were much higher than those for linear copolymers at similar composition

[133]. Similar improvement of both stress and elongation was also reported by Juhari and co-

workers on the study of poly(n-butyl acrylate)-b-poly(α-methylene-γ-butyrolactone) block

copolymers [259].

4.1.3 Polymers prepared by controlled radical polymerization

In contrast to ionic polymerization, controlled radical polymerizations have higher tolerance to

monomer functionalities and are suitable for both “core first” and “arm first” strategies. By using

a tri-functional ATRP initiator and “core first” strategy, three-arm star copolymers with PMMA

[260], polyacrylonitrile (PAN) [261], or PS[262] as a glassy segment and PnBA as elastic segment

were prepared for TPE properties evaluation. As an all acrylic TPE, three-arm star (PMMA-

PnBA)3 with 36 % of PMMA showed 11 MPa ultimate stress with 545 % strain at break. (PAN-

PnBA)3 star polymers displayed ultimate tensile strength from 6.3 MPa to 12.7 MPa as the strain

at break in the range from 382 % to 700 %. Phase separation between PAN and PnBA was retained

with service temperature up to 250 °C. As the temperature was further raised up to 280 °C, the
49

PAN domain started to crosslink chemically. The storage modulus of these materials dropped

when the temperature was close to 300 °C. With multifunctional ATRP initiators having 10 and

20 initiation sites, 10 arms and 20 arms PMBL/PnBA star polymers were prepared for high-

temperature TPE applications [259]. The highest ultimate tensile strength achieved was 7.8 MPa.

Strain at break was lower than 140 %.

4.2. Graft copolymers for TPEs

Generally, there are three strategies to synthesize graft copolymers. (1) In the “grafting onto”

method, pre-synthesized side chains are grafted onto the polymer backbone through reactions

between chain-end functionalities on side chains and in-chain functionalities on the backbone. (2)

In the “grafting from” method, monomers of the side chains are directly propagated from initiation

sites that are present on the polymer backbone. (3) In the “grafted through” or “macromonomer”

approach, side chains with polymerizable end groups are mixed with monomers used to grow the

backbone and polymerized together. The branched architectures of graft copolymers are typically

composed of a linear polymer backbone and have more than one polymeric side chain attached by

a covalent bond [13, 263]. The structure of graft copolymers, Figure 19, is defined by three

structural factors that ultimately allowed additional parameters to tailor the structure-property

relationship of the material: (1) the molecular weight of the main chain, (2) the molecular weight

of the grafted chains, and (3) the distance between the grafted chains [15].
50

Figure 19. Three structural factors of graft copolymers [15]. Copyright 2014, Reproduced

with permission from the American Chemical Society.

Ideally, well-defined graft copolymers consist of monodisperse side chains attached to a

monodisperse backbone with strict control over the both the number of branch points and the

regular spacing of the branch junctions, or so-called exact graft (eg) polymer. Most graft

copolymers, or randomly graft (rg) polymers, are described by an average number of graft branches

per molecules and random spacing distribution of graft branches along the backbone [13]. Figure

20, B-eg-A2 is the exactly graft copolymer with B as the backbone and A as the side-arm, where

2 of A blocks are grafted at each branch point. C-rg-(AB) is the randomly grafted polymers with

C as the backbone and AB diblock copolymer as the side-arm grafted from the end of B block.

Figure 20. Illustrations of exact graft and random graft polymers.

4.2.1. Multigraft copolymers type TPEs synthesized by living anionic polymerization

The grafting to, grafting from, and grafting through approaches have all been used to produce

polydiene-graft-PS materials. The grafting to strategy was accomplished by the hydrosilylation,

which has been reviewed elsewhere [264].


51

In order to qualify as exact multi-graft copolymers (MGC), the main chain, the graft chain, and

the distance between the graft chains need to be precisely synthesized with narrow polydispersity.

Both an iterative method and a chlorosilane anion coupling method have been developed based on

living anionic polymerization to prepare well-defined multi-graft copolymers. By using the

chlorosilane anion coupling method, Mays and co-workers synthesized a serious of regular spaced

trifunctional, tetrafunctional and hexafunctional MGC with PI as the elastic backbone and PS as

the glassy graft chains (Figure 20, graft polymers) [265-268]. These well-defined MGCs were

ideal materials to investigate mechanical properties of multi-graft copolymers, including chain

dynamics [269], tensile properties [270], deformation behavior [271, 272], stress softening [273]

and hysteresis behavior [274].

The tetrafunctional random multigraft copolymer with 23 vol% PS and 4.4 average grafting

points (RMG-4-23-4.4) showed a surprisingly high strain at break of about 1550%, which was far

exceeding the values for commercial TPEs (Styroflex®: 620%; Kraton®: 1050%) (Figure 21.a and

21.b) [274]. Even after elongating to 1400% strain, the exact tetrafunctional MGC with 14 vol%

PS and 5.5 average grafting points (MG-4-14-5.5) exhibits only 40% residual strain in hysteresis

experiments (Figure 21.d). Thus, these materials have become known as “superelastomers” for

their superior elongation properties and excellent elastic recovery [275]. To understand how the

branched architecture improves mechanical performance, consider the following. For a polymer

chain of ABA triblock copolymer, two glassy end blocks reside either in a single glassy domain to

form a loop or in different glassy domains to form a bridge that contributes to elastic recovery. In

contrast, the elastic backbone of MGCs are tethered to multiple glassy domains, thus improving

elastic recovery [271, 272]. The multiple tethering also improves load distribution throughout the

biphasic material. Besides the excellent elastic recovery and elongation properties, the tensile
52

stress of MGCs was decoupled from composition and tuned based on either branch points or

branch functionalities. For MGCs with about 20 vol% of PS and 4 branch points, as the tensile

stress increased with the number of branch functionalities, the morphology evolved from spherical

for trifunctional MGC, cylindrical for tetrafunctional MGC, to lamellar for hexafunctional MGC.

For tetrafunctional MGCs with about 20 vol% of PS, the tensile stress increased with the number

of branch points (Figure 21.c).

Figure 21. Mechanical properties of PI-g-PS multi-graft copolymers. (a) Influence of

functionality on stress-strain behavior. (b) Comparison of tetrafunctional multi-graft copolymers

(MG) with commercial SIS-block copolymer (Kraton®) and star copolymer (Styroflex®). (c)

Influence of the number of branch points on the tensile strength. (d) Hysteresis curve of a

tetrafunctional multi-graft copolymer [274]. Copyright 2006, Reproduced with permission from

John Wiley and Sons Inc.

4.2.2. Multigraft copolymers type TPEs synthesized by hybrid polymerization methods


53

The fundamental research on exact MGCs not only established the structure-property

relationship of these materials but also provided essential guidance to design MGCs for the

application as TPEs. However, synthesizing exact MGCs by the living anionic polymerization

method requires stringent oxygen and moisture free conditions, and multiple solvent/non-solvent

fractionations to isolate near-monodisperse products. In order to develop an economically viable

process, Zhang and Mays developed various hybrid polymerization by combining (mini)emulsion

polymerization with anionic polymerization or controlled radical polymerization to prepare

trifunctional and tetrafunctional grafted copolymers with PS or PMMA as side chains, with PI or

PnBA as the backbone [276-280].

In their work to synthesize all acrylic superelastomers, Mays and co-workers first prepared

PMMA macromonomers by living anionic polymerization using N-isopropyl-4-vinylbenzylamine

(PVBA) as the initiator in a one-batch process (Figure 22). The well-defined PMMA

macromonomers were randomly copolymerized with nBA by using RAFT polymerization. The

content of PMMA macromonomer in the final graft copolymers PnBA-rg-PMMA was extensively

suppressed. This straightforward and efficient method, as well as the increase of the molecular

weight and improvement of microphase separation behavior, lead to elongation at break around

1700% and strain recovery with less than 15% hysteresis [281].

Figure 22. Schematic illustrations of the sythesis of PMMA macromonomer by anionic

polymerization and the synthesis of PnBA-rg-PMMA by RAFT polymerization. Modified from

reference [281].
54

Fonagy, Tamas, and co-workers reported the synthesis of PIB-g-PS by combining cationic

polymerization with ATRP using the grafting from methodology. The initiation sites were

introduced into the polymer main chain by copolymerizing isobutylene with a trace amount of 4-

bromomethylstyrene. The bromomethyl group served as the ATRP initiation sites [282]. Suksawad,

Patjaree, and coworkers reported the preparation of graft copolymers with natural rubber as the

backbone [283]. After deproteinization, part of the double bonds in the deproteinized natural

rubber (DPNR) were attacked by tert-butyl hydroperoxide/tetraethylenepentamine to generate

initiation sites for the polymerization of styrene. The grafting efficiency of this method could

achieve more than 90 mol%. The resulting DPNR-g-PS exhibited four times higher ultimate stress

and two times higher ultimate strain than that of unmodified nature rubber. This stress-strain

behavior is similar to that of crosslinked nature rubber. Wei and Li reported the synthesis of star

graft TPEs by combining anionic polymerization with ROP. The rubbery backbone was prepared

by living anionic polymerization of myrcene followed by epoxidation and hydroxylation. Ring

opening polymerization of L-lactide was initiated from the hydroxyl group on the backbone. The

mechanical properties of resulting polymers were tunable based the grafting density and side chain

length [284].

Polyhedral oligomeric silsesquioxanes (POSS) are silicon-containing organic-inorganic

materials [285]. Hybrid materials with silicone combine thermal/oxidative resistance characteristic

of inorganic materials and processability of organic materials [286, 287]. Copolymerizing ethylene

and propylene with ethyl-POSS-norbornene by coordination polymerization produced random

terpolymers of ethylene-propylene-silsesquioxane [287]. High stress at break (>10 MPa) was

observed for these silica containing TPEs. The excellent crossover reactivity between olefin and
55

norbornene containing monomers was also used to copolymerize ethylene with 5-norbornene-2-

methylpropanoate (NBMP) which yielded a PE backbone with randomly distributed 2-bromo-2-

methylpropanoate as ATRP initiation sites [288]. Grafting nBA by ATRP produced PE-g-PnBA

TPEs with 15 MPa ultimate stress and 490 % strain at break. The same group further prepared PE-

b-(PE-co-PNBMP)-b-PE, a triblock copolymer with PE at both end and macroinitiator in the

middle block [289]. Graft polymers showed an ultimate tensile strength of 27 MPa with 1310 %

strain at break.

4.2.3. Multigraft copolymer type TPEs based on rigid backbones

Traditionally there are two types of macromolecular architectures that form a continuous

rubbery matrix and are thus suitable for TPEs: (1) linear and star polymers with glassy end (outer)

blocks and elastic middle (inner) block (architecture I); (2) graft polymers with elastic backbone

and glassy or semicrystalline side chains (architecture II). Tang and Wang proposed a reversed

graft architecture with elastic side chains grafted on rigid backbone (architecture III) [290]. Rigid

polymers and inorganic materials such as cellulose [290-294], carbon nanotubes [295, 296], lignin

[297, 298], silica nanoparticles [299], and iron magnetic particles [300] could be used as the

backbone. Generally, the “grafting from” approach is used to prepare TPEs based on rigid

backbones.

Cellulose contains multiple hydroxyl groups along the chain. By reacting these hydroxyl groups

with 2-bromoisobutyryl bromide, Jiang and co-workers prepared ATRP-initiator functionalized

cellulose to copolymerize MMA and nBA. (Figure 23.a) [290]. Compared with linear, star and

graft all acrylic TPEs, graft copolymers with the reversed architecture displayed higher ultimate

stress (11.1 MPa) with modest strain at break (550 %). One reason for the high stress was attributed

to the hydrogen bonding between the carbonyl group on acrylic side chains and hydroxyl groups
56

on cellulose. Microphase separation was observed between all acrylic side chains and cellulose

backbone. Sustainable TPEs with monomers derived from rosin and fatty acids were prepared by

using the same strategy [292]. Zhang and Hillmyer directly polymerized hydroxypropyl

methylcellulose (HPMC) from the hydroxyl groups on cellulose by ring opening polymerization

and prepared poly(β-methyl-δ-valerolactone)-b-poly(L-lactide) (PMVL-b-PLLA) copolymers

(Figure 23.b) [301]. These materials are considered to be thermoplastics.

Figure 23. Graft from methodology to make graft copolymer TPEs based on cellulose matrix

through (a) ATRP [290]; Copyright 2013, Reproduced with permission American Chemical

Society. (b) ring opening polymerization [301]. Copyright 2016, Reproduced with permission

from the American Chemical Society.

Grafting polyisoprene onto the backbone of cellulose was achieved by Wang [293, 294], and the

resulting materials could mimic mechanical properties of human skin [293]. Different from

cellulose-g-(PnBA-co-PMMA), for which only one Tg corresponding to (PnBA-co-PMMA) was

reported [293], two distinguishable Tgs were observed through dynamic mechanical analysis in

cellulose-g-PI. By replacing cellulose with functionalized carbon nanotubes (CNT) [295] and

Fe3O4 nanoparticles (NPs) [300], TPEs with architecture III were successfully prepared. When
57

carbon nanotubes were used as the rigid segment, 1.7 % of CNT loading increased the ultimate

tensile stress from 2.4 MPa to 7.8 MPa. Strain at break was reduced from 951 % to 520 %. TPEs

with Fe3O4 showed recyclability under a magnetic field.

4.3. Bottlebrush Copolymers for TPEs

Bottlebrush polymers are highly branched polymers with tightly spaced side-chains grafted on

the polymer backbone, ultimately with one side chain at each backbone repeat unit [302]. Typically,

the highly extended cylindrical conformation prevented intermolecular entanglement and led to

Rouse-like relaxation without observing a rubbery plateau [303-305]. Pakula and Matyjaszewski

first prepared either Abb-Bbb-Abb triblock or (Abb-Bbb)3-star bottlebrush polymers (Figure 24)

containing poly(octadecyl methacrylate) (PODMA) as the hard block A and PnBA bottlebrush as

the soft segment B [306]. These materials demonstrated supersoft properties with an equilibrium

modulus of 1 kPa, which is typically observed in hydrogels. In contrast with hydrogels which

required solvent (water) to swell the polymer network, the bottlebrush polymers were swollen by

the dangling side chains which act as permanent diluent.

Figure 24. Schematic illustrations of block and star bottlebrush polymers for TPE applications.

The low modulus characteristic of bottlebrush polymers especially fit the application

requirements in tissue engineering or medical implants [307], where soft elastomers with modulus

as low as 102 to 103 Pa are preferred (Figure 25. a and b) [308]. In comparison, conventional linear
58

polymers possess an intrinsic entanglement modulus ranging from 105 to 106 Pa. The solvent-free

system with bottlebrush supersoft elastomers avoids issues associated with hydrogels, such as

sensitivity to mechanical deformation and solvent depletion upon compression. With the incentive

of developing solvent free supersoft elastomers for biological tissues [308], Sheiko evaluated the

structural influence of three parameters - degree of polymerization of side-chain (nsc), grafting

density (ng), and degree of crosslinking (nx) (Figure 25.c). Mimicking biological stress-strain

behavior with synthetic bottlebrush polymers was achieved through a three-step fitting, modeling,

and synthesis process (Figure 25.d). Linear-bottlebrush-linear ABA triblock type polymers were

designed to mimic skin tissue with the feature of adaptive coloration [309].

Figure 25. a) The mechanical diversity of biological and synthetic materials. b) Inverse

relationship between elongation at break and Young’s modulus of different materials. c) structural

parameters: degree of polymerization of side-chain (nsc), grafting density (ng), and degree of

crosslinking (nx). d) Flowchart describing how to replicate specific mechanical properties of

biological tissues in brush-like elastomers [308]. Copyright 2017, Reproduced with permission

from Springer Nature.


59

4.4 Polymers with other architectures

Hyperbranched polymers are a type of dendritic polymer with high branching density and

randomly distributed branches [310]. Typically synthesized in one-pot processes, the synthetically

facile method leads to a high level of structural irregularities and broad polydispersity in

hyperbranched polymers (Figure 26). By using living anionic polymerization, Hutchings prepared

an α,ω,ω`-trifunctional PS-PI-PS triblock macromonomer. Subsequent Williamson etherification

between the α- bromo group and ω,ω`-dihydroxyl group condensed macromonomers into

hyperbranched block copolymers [311, 312]. Despite their structural irregularity and broad

molecular weight distribution, the resulting hyperbranched PS-PI copolymers displayed 1643%

strain at break and 12.4 MPa tensile strength. Blending 10% hyperbranched copolymers with

Kraton D-1160 improved the mechanical properties. The morphological characterization indicated

that the hyperbranched polymers lacked long-range order due to the highly branched architecture.

[313] Nugay and Kennedy recently reported a low cost bifunctional cationic polymerization

initiator to prepare PIB and PS-PIB-PS triblock copolymers. During the synthesis of the triblock

copolymer, a significant amount of high molecular weight hyperbranched copolymers was

discovered after the complete consumption of styrene. While the tensile strength and elongation

remained similar, the blend of triblock and hyperbranched copolymers exhibited lower permanent

set and deformation under load, and higher yield strength and toughness than pure triblock

copolymer [314].
60

Figure 26. Schematic illustrations of hyperbranched polymer and POM-POM polymer

architectures.

AnBAn, often referred as dumbbell [315] or pom-pom [316] copolymers, are star-linear-star

triblock copolymers with An star polymer as the end block (Figure 26). Huang and Knauss first

synthesized well-defined pom-pom copolymers by living anionic polymerization in one-pot with

PI, PS or PDMS as the linear middle block and (PS)n as the star-shaped end block [316-318].

Morphological analysis indicated that the increase of the length of the middle block and the number

of arms frustrated the formation of ordered cylindrical phases and promoted the transition to a

disordered phase [319]. However, the detailed mechanical properties were not presented. Hirao

and Faust prepared (PMMA)n-PIB-(PMMA)n pom-poms by a combination of living anionic and

cationic polymerization [320]. The number of PMMA arms was varied as 2, 4 and 8. All polymers

showed fair to good mechanical properties with 12.6 - 20.7 MPa tensile strength and 279% - 444%

strain at break. However, the Mn of both PIB and PMMA was designed to maintain a constant

overall molecular weight. Thus, the effects of the number of arms on the mechanical properties

remain unclear. Kamigaito prepared pom-pom copolymers by ruthenium-catalyzed living radical

polymerization [321]. With PS as the star block and poly(dodecyl methacrylate) (PDMA, Tg ~50
oC) as the linear block, a cylindrical morphology was observed with 47 wt% PS fraction. Over the
61

range of 17 ~ 30 wt% of PS, the materials exhibited tensile strengths of 1 ~ 6 MPa and elongations

at break of 70 ~ 300% [322].

5. TPEs based on supramolecular chemistry

Traditional TPEs rely on thermodynamic incompatibility between different segments to form

physically crosslinked networks. Supramolecular interactions introduce additional interactions

into polymer systems with reversible dynamic networks. TPE materials with self-healing and/or

shape memory properties have been prepared by fusing block copolymer systems with

supramolecular interactions, such as ion interaction, hydrogen bonding, π-π stacking and metal-

ligand coordination.

5.1 TPEs with hydrogen bonding

Since the ‘kick-start’ of supramolecular polymer chemistry in 1997 with 2-ureido-4[1H]-

pyrimidinone (UPy) [323], various hydrogen bonds donor/acceptor pairs have been incorporated

into elastic polymers, either on the main-chain [324, 325] or on chain ends [326, 327], to prepare

supramolecular elastomers where dimensional stability relied entirely on noncovalent interactions.

Leibler first explored the concept of preparing self-healing rubbery materials from supramolecular

assembly through hydrogen bonding. By reacting fatty acid dimer and trimer with diethylene

triamine and urea, a rubber-like oligomer mixture with complementary hydrogen bonding was

prepared with strain at break over 500% and ultimate stress higher than 2.5 MPa (Figure 27) [328].

This breakthrough not only prepared a self-healing rubber, but also provide insight to fuse

hydrogen bonds into block copolymer to promote dynamic secondary interactions.


62

Figure 27. Synthesis pathway of self-healing and thermoreversible rubber from hydrogen

bonding. The hydrogen bond donors are shown in green, the acceptors in green [328]. Copyright

2008, Reproduced with permission from Springer Nature.

In TPE materials, hydrogen bonds donors/acceptors can be integrated either on hard/soft

segments or on both segments. Long prepared nucleobase functionalized triblock copolymer by

NMP with polystyrene derivatives containing either adenine (Ade-) or thymine (Thy-) in the end

block. Complementary blending Ade- and Thy- containing triblock copolymers formed a single

hard phase and significantly increased the solution viscosity [329]. Blending Ade- containing

polymers with uracil-containing phosphonium salt softened the hard phase but improved the

storage modulus [330]. Using RAFT polymerization, Noro prepared triblock copolymers with poly

(n-butyl acrylate)-co-polyacrylamide (Ba) as the middle elastic block and P4VP [331] or PS [332]

as the end hard segments (Figure 28). In both case, polyacrylamide served as the multi hydrogen

bond donor to dynamically crosslink the middle block. Additionally, high-pressure RAFT

polymerization protocol was utilized to obtain poly(n-butyl acrylate)-co-polyacrylamide with

molecular weight over 1000 kg/mol. Guan designed a grafted copolymer with PS as the rigid

backbone and poly(acrylate amide) as the side chain [333]. The materials displayed self-healing
63

properties with high modulus and toughness. Since the molecular weight of poly (acrylate amide)

was lower than the entanglement molecular weight, the elasticity was mainly from hydrogen

bonding in the poly(acrylate amide) domains.

Figure 28. a) Chemical structure and schematic representation of S-Ba-S. Photos and schematic

networks of S-Ba-S elastomers b) before elongation and c) after elongation are also shown [332].

Copyright 2016, Reproduced with permission from John Wiley and Sons Inc.

5.2 TPEs with ionic interactions

Ionic groups, historically, have been incorporated into hydrocarbon polymers to improve their

thermal properties and enhance their compatibility with polar additives. Another benefit of

introducing ionic bonds into polymers is to facilitate their potential application in flexible

electronics and ion exchange membranes due to their ion conductivity. Long prepared ABA

triblock copolymer through NMP with the phosphonium ion in A blocks and PnBA as B block

[334]. The storage modulus increased as the mole percentage of ions was increased. The presence

of ionic content strongly shifted the phase boundary and resulted in a lamellar phase with 21% of

the phosphonium domain. Hoogenboom prepared a self-healing TPEs by blending oppositely


64

charged oligomeric ABA triblock copolymer with charge in the A blocks [335]. These materials

were able to fully recover their storage modulus within 2 hours upon mild heating. Wang prepared

ABA triblock copolymers with 2% of 1-vinylimidazole as the co-monomer in B block and

evaluated the influence of different level of ionic crosslinking in B block on the mechanical

properties. Since imidazole groups were randomly distributed in the B block, crosslinking with

1,6-dibromohexane reduced the ultimate strain (Figure 29) [336]. However, tensile stress and

elastic recovery were significantly enhanced. The same group prepared BAB triblock copolymers

with imidazole containing (2 wt%) elastic B block as the end blocks [337]. In these materials with

reversed block sequence, typical TPS stress-strain behavior was observed, which might be

attributed to the presence of ionic association in the B block.

Figure 29. Illustration of Synthesis of (a) 1-vinylimidazole containing triblock copolymers and

(b) Ionically cross-linked TPEs [336]. Copyright 2017, Reproduced with permission from the

American Chemical Society

5.3 TPEs with π-π stacking

Arene-arene stacking or π-π stacking promotes noncovalent interactions between aromatic rings

due to the great binding affinity between π bonds. Many researchers have taken advantage of π-π
65

stacking between polystyrene and aromatic containing compounds, such as benzoxazine, carbon

black, carbon nanotubes, graphene, and graphene oxide, to enhance the mechanical performance

of styrenic elastomers and rubbers [338-342]. Burattini and coworkers prepared self-healing

elastomers by blending a chain-folding polyimide and a pyrenyl telechelic functionalized

polyurethane. The π-electron-deficient bis(diimide) motif exhibits a strong encapsulation with π-

electron-rich pyrenyl groups that combines π-π stacking with complementary electronic

interactions into one system [343, 344]

5.4 TPEs with metal-ligand coordination.

Metal-ligand coordination is another supramolecular interaction that can be incorporated into

TPEs systems. Mozhdehi reported a polymer composed of PS as the hard domain and PnBA as

the soft domain [345]. In addition, imidazole containing acrylate monomer was copolymerized

with the nBA monomer to introduce zinc-imidazole interactions. Tunable stress-strain behavior

was achieved with the change of backbone composition, imidazole containing acrylate content,

and ligand/Zn ratio. The polymer showed excellent self-healing properties that were more tolerant

to moisture and have broader range of tunability of thermodynamics and kinetics than that of

polymers based on hydrogen bonding interactions (Figure 30). Similarly, TPEs with poly (4-vinyl

pyridine) (P4VP) as the hard segment, such as P4VP-b-PB-b-P4VP triblock copolymers, also

exhibited great improvement of modulus, tensile strength, and toughness through the interaction

with Zn2+ [346]. The dissociation/reformation of metal-ligand bonds by the temperature control

also endowed such polymers with shape memory behavior [347]. Sequentially blending carboxyl-

terminated polybutadiene with zinc acetate dihydrate and poly(styrene-co-4vinylpyridine) also

resulted in supramolecular TPEs with shape memory behavior [348]. Multidentate ligands have

higher binding constants and stronger interaction with metal ions. Ligands such as 2,6-bis(19-
66

methylbenzimidazolyl) pyridine (Mebip) [349], 4-oxy-2,6-bis- (10-methylbenz imidazolyl)

pyridine (HOMebip) [350] and 1,3-benzenedicarboxamide [351] were successfully integrated into

polymer network to prepare stimulus responsive supramolecular polymers. The resulting polymers

all showed excellent seal-healing performance. However, these polymers often contain high

concentration of metals which could limit certain applications.

Figure 30. Schematic illustrations of metal-ligand coordination interactions between Zn2+ and

imidazole [345]. Copyright 2014, Reproduced with permission from the American Chemical

Society.

6. Conclusion and outlook:

Various thermoplastic elastomers have been developed over the course of the past six decades.

The global value of thermoplastic elastomers is projected to grow from USD 20.88 billion in 2016

to USD 28.27 USD billion by 2022 [352]. Advancement in synthetic chemistry is the key driving

force that facilitates the successful transition of TPEs from inventions in the laboratory to

commodities in the marketplace. Along with the progress that has been made, challenges still

remain in the field of synthetic polymer chemistry. SBCs prepared by living anionic

polymerization are expected to lead the market of TPEs over the next decade. However, all SBCs
67

suffer from limited thermal resistance due to the relatively low glass transition temperature of PS.

The key to increase the upper service temperature of SBCs at the industrial level is to find a

monomer that is able to copolymerize anionically with dienes in hydrocarbon solvents at mild

temperatures. In this way, the existing manufacturing process can remain largely unaffected. PBF

might serve as an interesting alternative to PS for high temperature application [109]. However,

preparing PBF-PI-PBF triblock copolymers requires the difunctional lithium initiator approach. In

addition, the synthesis of benzofulvene monomer is relatively challenging with low yields and low

monomer stability on storage. With the Lewis base and di-phenoxyalkyl aluminum initiation

system (LA system), Kuraray introduced all acrylic TPEs with better UV resistance and greater

transparency as compared to SBCs [99]. One benefit of the LA system is its capacity to prepare

well-defined PMMA-PnBA-PMMA triblock copolymers at economically viable polymerization

temperatures. Exploring additional methacrylate and acrylate derivatives with LA systems is

worthwhile in order to prepare functional TPEs for specific applications.

The limitations of applying controlled radical polymerizations, such as ATRP and RAFT

polymerization, include higher cost of controlling agents, additional cost to eliminate metal or thio-

containing end groups, and macroinitiator purification. However, controlled radical

polymerizations, provide access to a wide range of functional monomers. For ATRP, the difficulty

in quantitively removing metal catalysts is a concern for material applications in food packing or

medical devices. Recent progress in developing new catalytic systems [353] and photoinduced

electron transfer reactions [354], as well as (mini)emulsion polymerization [355] and

polymerization reactors [356], could reduce the amount of catalysts and accelerate the

polymerization rate, which might provide novel practical applications of ATRP. Other than the

ability to incorporate various functional monomers, controlled radical polymerization techniques,


68

especially RAFT polymerization, provide a more robust process to prepare polymers with

advanced macromolecular architectures such as Arkema’s all acrylic block copolymers [33] and

Lubrizol’s star-shaped polymer viscosity modifier [357], which have achieved industrial level

applications. More applications of controlled radical polymerization are summarized in a recent

review [358].

As the production and consumption of synthetic polymers increased exponentially over the past

50 years, many economic and environmental issues associated with polymers from petroleum

resources have attracted the attention of the public and directed polymer chemists to re-think the

feedstocks for future plastics by employing sustainable resources, without jeopardizing food

supplies [359, 360]. Tremendous progress has been made in using ROP to prepare thermoplastic

elastomers from sustainable materials. As pointed out by Hillmyer, a comprehensive

understanding of both synthetic limitations and block copolymer structure-property relationships

will be the key to designing biodegradable plastics from sustainable resources [152]. This effort

will also require a collaboration among biologists, biochemists, and polymer chemists to prepare

monomers and polymers from widely accessible materials.

Most of the current TPEs are linear block copolymers. Branched macromolecular architectures,

especially miktoarm star and multi-graft copolymers, deliver additional controlling parameters

with which to decouple the morphological dependence on polymer composition. Consequently,

TPEs with previously inaccessible stiffness and elasticity were prepared. Branched polymers with

bottlebrush architectures can achieve stiffnesses that are similar to those of biological organs.

Unlike hydrogels, which require water to swell the polymer network, bottlebrush copolymers

utilize the densely grafted side-chain to swell the polymer network, and these systems demonstrate

supersoft properties without the presence of liquid solvent. However, polymers with complicated
69

architectures are challenging to prepare in industry. Towards this goal, a series of PnBA-rg-PS2

multi-graft copolymers has been prepared by a combination of anionic and emulsion

polymerization [276]. The practical drawback of this method is that most factories do not have

both solution and emulsion polymerization facilities on the same site. To truly benefit from the

unexpected but remarkable properties that are endowed by star, multi-graft and bottlebrush

polymers, more practical approaches need to be established to prepare these polymers in

conventional polymer manufacturing facilities on a commercial scale.

Along with the progress made in polymerization methodologies and control of macromolecular

architecture, achievements in click chemistry and supramolecular chemistry have ungraded the

toolbox of polymer chemists to allow preparation of functional TPEs with self-healing and

stimulus responsive abilities. For industrial applications of TPEs as either commodity or specialty

materials, a balance between the cost and the performance of materials needs to be taken into

consideration. This balance could be achieved by smart choice of monomers, polymerization

methodologies, macromolecular architectures, and post-polymerization modification, and

potentially expanding the application of TPEs from traditional fields, such as automotive,

footwear, adhesives, and food packaging, to 3D-printing, flexible electronics, separation

membranes and tissue engineering.

Acknowledgments

This work was supported by the U.S. Department of Energy, Office of Science, Basic Energy
Science, Materials Sciences and Engineering Division. A portion of the synthesis and
characterizations was conducted at the Center for Nanophase Materials Sciences, which is a DOE
Office of Science User Facility WW and PY are grateful to the support of the National Key
Research and Development Program of China (No. 2018YFB0704200) and National Thousand
Talent Plan. We sincerely thank the reviewers of this manuscript for their constructive criticisms
which helped us to improve this review.
70

Reference

[1] Bonart R. Thermoplastic elastomers. Polymer 1979;20:1389-403.


[2] Spontak RJ, Patel NP. Thermoplastic elastomers: fundamentals and applications. Curr
Opin Colloid Interface Sci 2000;5:333-40.
[3] Shanks RA. General purpose elastomers: structure, chemistry, physics and performance.
In: Visakh P.M, Thomas S, Chandra A.K, Mathew A.P, editors. Advances in Elastomers
I. Heidelberg: Springer, 2013. p. 11-45.
[4] Walker BM, Rader CP. Handbook of Thermoplastic Elastomers. New York: Van
Nostrand Reinhold, 1979. 430 pp.
[5] Bhowmick AK, Stephens H. Handbook of elastomers. New York: CRC Press, 2000.
944 pp.
[6] Fakirov S. Handbook of condensation thermoplastic elastomers. New York: John Wiley
& Sons Inc; 2006. 643 pp.
[7] Drobny JG. Handbook of thermoplastic elastomers. Amsterdam: Elsevier Ltd, 2014.
736 pp.
[8] Bayer O, Siefken W, Rinke H, Orthner L, Schild H. A process for the production of
polyurethanes and polyureas. DRP 728981, 1937.
[9] Matuszak MP. Production of isooctanes from cyclopropane and isobutane. US 2632031
A, 1953.
[10] Langerak EO, Prucino LJ, Remington WR. Elastomers from polyalkylene ether glycol
reacted with arylene diisocyanate and water. US 2692873 A, 1954.
[11] Schollenberger CS. Simulated vulcanizates of polyurethane elastomers. US 2871218 A,
1959.
[12] Geoffrey H, Ralph M. Block polymers of monovinyl aromatic hydrocarbons and
conjugated dienes. US 3265765 A, 1966.
[13] Hsieh H, Quirk RP. Anionic Polymerization: Principles and Practical Applications.
Boca Raton: CRC Press; 1996. 744 pp.
[14] Handlin DL, Trenor S, Wright K. Applications of thermoplastic elastomers based on
styrenic block copolymers. In: Matyjaszewski K, Gnanou Y, Leibler L, editors.
71

Macromolecular Engineering: Precise Synthesis, Materials Properties, Applications.


New York: John Wiley & Sons, 2007. p. 2001-31.
[15] Hirao A, Goseki R, Ishizone T. Advances in Living Anionic Polymerization: From
Functional Monomers, Polymerization Systems, to Macromolecular Architectures.
Macromolecules 2014;47:1883-905.
[16] Szwarc M. Living' polymers. Nature 1956;178:1168-9.
[17] Orlov Y. Polymer thermodynamics: Leverage modeling results for industrial
applications. In: Vivaldo-Lima E, Debling J, Zaldo-Garcia F, Tsavalas J, editors. Proc
Polymer Reaction Engineering IX 2015. Abstract No. 28.
[18] Drolet F, Fredrickson GH. Combinatorial screening of complex block copolymer
assembly with self-consistent field theory. Phys Rev Lett 1999;83:4317-20.
[19] Teraoka I. Polymer solutions: An introduction to physical properties. New York: John
Wiley & Sons Inc; 2002. 360 pp.
[20] Quirk RP. Applications of Anionic Polymerization Research. Washington DC: ACS
Publications, 1998. 332 pp.
[21] Holden G. Thermoplastic Elastomers. New York: John Wiley & Sons Inc; 2000. 618
pp.
[22] Ahuja K, Deb S. Thermoplastic Elastomers (TPE) Market Size By Product, Industry
Analysis Report, Regional Outlook, Application Potential, Price Trend, Competitive
Market Share & Forecast, 2016-2023. https://www.gminsights.com/industry-
analysis/thermoplastic-elastomers-tpe-market-report 2018;Accessed Sep 2018.
[23] Richard H, Karl M. Composition of matter comprising polypropylene and an ethylene-
propylene copolymer. US 3262992 A, 1966.
[24] Coran AY, Patel RP. Elastoplastic compositions of butyl rubber and polyolefin resin.
US 4130534 A, 1978.
[25] Witsiepe WK. Segmented thermoplastic copolyester elastomers. US 3651014 A, 1972.
[26] Chen AT, Farrissey WJ, Robert JGNI. Polyester amides suitable for injection molding.
US 4129715 A, 1978.
[27] Hadjichristidis N, Hirao A, editors. Anionic Polymerization: Principles, Practice,
Strength, Consequences and Applications. Tokyo: Springer Japan; 2015. 1079 pp.
72

[28] Aoshima S, Kanaoka S. A renaissance in living cationic polymerization. Chem Rev


2009;109:5245-87.
[29] Khosravi E, Szymanska-Buzar T. Ring opening metathesis polymerisation and related
chemistry: state of the art and visions for the new century. Netherlands: Springer
Science & Business Media; 2012. 493 pp.
[30] Matyjaszewski K, Tsarevsky NV. Macromolecular engineering by atom transfer radical
polymerization. J Am Chem Soc 2014;136:6513-33.
[31] Moad G, Rizzardo E, Thang SH. Radical addition–fragmentation chemistry in polymer
synthesis. Polymer 2008;49:1079-131.
[32] Hawker CJ, Bosman AW, Harth E. New polymer synthesis by nitroxide mediated living
radical polymerizations. Chem Rev 2001;101:3661-88.
[33] Boutillier JM, Disson JP, Havel M, Inoubli R, Magnet S, Laurichesse , et al. Self-
assembling acrylic block copolymers for enhanced adhesives properties.
https://www.adhesivesmag.com/articles/91909-self-assembling-acrylic-block-
copolymers-for-enhanced-adhesives-properties 2013;Accessed Sep 2018.
[34] Ariura F. Acrylic block copolymer for adhesive application. J Adhes Soc Jpn
2013;49:336-42.
[35] Seymour RB, Kauffman GB. Elastomers: III. thermoplastic elastomers. J Chem Educ
1992;69:967-70.
[36] Bates FS, Fredrickson GH. Block copolymers-designer soft materials. Phys Today
2000;52:32-8.
[37] Ober CK. Directed self-assembly: A dress code for block copolymers. Nat Nanotechnol
2017;12:507-8.
[38] Sinturel C, Bates FS, Hillmyer MA. High χ–low N block polymers: how far can we go?
ACS Macro Lett 2015;4:1044-50.
[39] Fetters LJ, Firer EM, Dafauti M. Synthesis and properties of block copolymers. 4.
poly(p-tert-butylstyrene-diene-p-tert-butylstyrene) and poly(p-tert-butylstyrene-
isoprene-styrene). Macromolecules 1977;10:1200-7.
[40] Kobayashi S, Kataoka H, Ishizone T, Kato T, Ono T, Kobukata S, et al. Synthesis and
properties of new thermoplastic elastomers containing poly[4-(1-adamantyl)styrene]
hard segments. Macromolecules 2008;41:5502-8.
73

[41] Bates FS, Hillmyer MA, Lodge TP, Bates CM, Delaney KT, Fredrickson GH.
Multiblock polymers: Panacea or pandora’s box? Science 2012;336:434-40.
[42] Misichronis K, Wang W, Cheng S, Wang Y, Shresthaa U, Dadmun M, et al. Design,
synthesis, and characterization of lightly sulfonated multigraft acrylate-based copolymer
superelastomers. RSC Adv 2018;8:5090-8.
[43] Sakurai S, Hashimoto T, Fetters LJ. Morphology of polystyrene-block-poly (ethylene-
alt-propylene) diblock copolymers in the strong-segregation limit. Macromolecules
1995;28:7947-9.
[44] Morton M, McGrath JE, Juliano PC. Structure-property relationships for styrene-diene
thermoplastic elastomers. J Polym Sci Part C Polym Symp 1969;26:99-115.
[45] Holden G, Bishop ET, Legge NR. Thermoplastic elastomers. J Polym Sci Part C Polym
Symp 1969;26:37-57.
[46] Khandpur AK, Foerster S, Bates FS, Hamley IW, Ryan AJ, Bras W, et al. Polyisoprene-
polystyrene diblock copolymer phase diagram near the order-disorder transition.
Macromolecules 1995;28:8796-806.
[47] Kim G, Libera M. Morphological development in solvent-cast
polystyrene−polybutadiene−polystyrene (sbs) triblock copolymer thin films.
Macromolecules 1998;31:2569-77.
[48] Fetters LJ, Morton M. Synthesis and properties of block polymers. I. poly-α-
methylstyrene-polyisoprene-poly-α-methylstyrene. Macromolecules 1969;2:453-8.
[49] Bolton JM, Hillmyer MA, Hoye TR. Sustainable thermoplastic elastomers from
terpene-derived monomers. ACS Macro Lett 2014;3:717-20.
[50] Kobayashi S, Matsuzawa T, Matsuoka SI, Tajima H, Ishizone T. Living anionic
polymerizations of 4-(1-adamantyl) styrene and 3-(4-vinylphenyl)-1, 1 ‘-biadamantane.
Macromolecules 2006;39:5979-86.
[51] McCormick HW. Ceiling temperature of α‐methylstyrene. J Polym Sci 1957;25:488-
90.
[52] Worsfold DJ, Bywater S. Anionic polymerization of α‐methylstyrene. J Polym Sci
1957;26:299-304.
[53] Shingo K, Kataoka H, Ishizone T. Living anionic polymerization of styrenes containing
adamantyl skeletons. J Physics Confer Ser 2009;184:012017/1-7.
74

[54] Kang BG, Shoji H, Kataoka H, Kurashima R, Lee JS, Ishizone T. Living anionic

polymerization of n‑(1-adamantyl)‑n‑4-vinylbenzylideneamine and n‑(2-adamantyl)‑n‑

4-vinylbenzylideneamine: Effects of adamantyl groups on polymerization behaviors and


thermal properties. Macromolecules 2015;48:8489-96.
[55] Matsuoka D, Goseki R, Uchida S, Ishizone T. Living anionic polymerization of 1-
adamantyl 4-vinylphenyl ketone. Macromol Chem Phys 2017;218:1700015/1-6.
[56] Kobayashi S, Kataoka H, Goseki R, Ishizone T. Living anionic polymerization of 4-(1-
adamantyl)-α-methylstyrene. Macromol Chem Phys 2018;219:1700450/1-7.
[57] Kobayashi S, Kataoka H, Ishizone T. Synthesis of well-defined poly(ethylene-alt-1-
vinyladamantane) via living anionic polymerization of 2-(1-adamantyl)-1,3-butadiene,
followed by hydrogenation. Macromolecules 2009;42:5017-26.
[58] Kobayashi S, Kataoka H, Ishizone T, Kato T, Ono T, Kobukata S, et al. Synthesis of
well-defined random and block copolymers of 2-(1-adamantyl)-1,3-butadiene with
isoprene via anionic polymerization. React Funct Polym 2009;69:409-15.
[59] Ishizone T, Tajima H, Torimae H, Nakahama S. Anionic polymerizations of 1-
adamantyl methacrylate and 3-methacryloyloxy-1,1'-biadamantane. Macromol Chem
Phys 2002;203:2375-84.
[60] Fuchise K, Sone M, Miura Y, Sakai R, Narumi A, Sato SI, et al. Precise synthesis of
poly(1-adamantyl methacrylate) by atom transfer radical polymerization. Polym J
2010;42:626-31.
[61] Lu W, Huang C, Hong K, Kang NG, Mays JW. Poly(1-adamantyl acrylate): Living
anionic polymerization, block copolymerization, and thermal properties.
Macromolecules 2016;49:9406-14.
[62] Yuki H, Hotta J, Okamoto Y, Murahashi S. Anionic copolymerization of styrene and
1,1-diphenylethylene. Bull Chem Soc Jpn 1967;40;2659-63.
[63] Xu JJ, Bates FS. Synthesis and thermal properties of hydrogenated poly (styrene-co-1,
1-diphenylethylene) copolymers. Macromolecules 2003;36:5432-4.
[64] Hutchings LR, Brooks PP, Parker D, Mosely JA, Sevinc S. Monomer Sequence Control
via Living Anionic Copolymerization: Synthesis of Alternating, Statistical, and
Telechelic Copolymers and Sequence Analysis by MALDI-TOF Mass Spectrometry.
Macromolecules 2015;48:610-28.
75

[65] Hutchings LR, Brooks PP, Shaw P, Ross-Gardner P. Fire and Forget! One-Shot
Synthesis and Characterization of Block-Like Statistical Terpolymers via Living
Anionic Polymerization. J Polym Sci Part A Polym Chem 2019;57:382-94.
[66] Yu JM, Dubois P, Teyssie´ P, Jerome R. Syndiotactic poly(methyl methacrylate)
(sPMMA)-polybutadiene (PBD)-sPMMA triblock copolymers: Synthesis, morphology,
and mechanical properties. Macromolecules 1996;29:6090-9.
[67] Schomaker E, Challa G. Complexation of stereoregular poly(methy1 methacrylates).
11. a mechanistic model for stereocomplexation in the bulk. Macromolecules
1988;21:2195-203.
[68] Schomaker E, Challa G. Complexation of stereoregular poly(methy1 methacrylates).
13. Influence of chain length on the process of complexation in dilute solution.
Macromolecules 1988;21:3506-10.
[69] Yu JM, Yu YS, Dubois P, Jerome R. Stereocomplexation of sPMMA-PBD-sPMMA
triblock copolymers with isotactic PMMA: 1. Thermal and mechanical properties of
stereocomplexes. Polymer 1997;38:2143-54.
[70] Yu J, Jérôme R. Stereocomplexation of sPMMA–PBD–sPMMA triblock copolymers
with isotactic PMMA. II: Effect of molecular weight. Polymer 1998;39:6567-75.
[71] Yu JM, Dubois P, Jérôme R. Poly [alkyl methacrylate-b-butadiene-b-alkyl
methacrylate] triblock copolymers: Synthesis, morphology, and mechanical properties at
high temperatures. Macromolecules 1996;29:8362-70.
[72] Hadjichristidis N, Mays J, Ferry W, Fetters LJ. Properties and chain flexibility of poly
(di-isobornyl methacrylate). J Polym Sci Polym Phys Ed 1984;22:1745-51.
[73] Yu JM, Dubois P, Jerome R. Synthesis and properties of poly[isobornyl methacrylate
(IBMA)-b-butadiene (BD)-b-IBMA] copolymers: New thermoplastic elastomers of a
large service temperature range. Macromolecules 1996;29:7316-22.
[74] Yu JM, Dubois P, Jerome R. Poly[poly(isobornyl methacrylate-co-methyl methacrylate)
(poly(IBMA-co-MMA))-b-polybutadiene-b-poly(IBMA-co-MMA)] Copolymers:
Synthesis, Morphology, and Properties. Macromolecules 1997;30:6536-43.
[75] Lu W, Yin P, Osa M, Wang W, Kang NG, Hong K, et al. Solution properties,
unperturbed dimensions, and chain flexibility of poly(1-adamantyl acrylate). J Polym
Sci Part B Polym Phys 2017;55:1526-31.
76

[76] Nakano Y, Sato E, Matsumoto A. Synthesis and thermal, optical, and mechanical
properties of sequence-controlled poly(1-adamantyl acrylate)-block-poly(n-butyl
acrylate) containing polar side group. J Polym Sci Part A Polym Chem 2014;52:2899-
910.
[77] Baskaran D, Müller AH. Anionic vinyl polymerization-50 years after Michael Szwarc.
Prog Polym Sci 2007;32:173-219.
[78] Foster FC. Butadiene polymers prepared with a lithium based catalyst characterized by
having at least 29 percent of the butadiene as cis-1, 4. US 3317918 A, 1967.
[79] Bywater S, Worsfold D. Anionic polymerization of isoprene. Ion and ion-pair
contributions to polymerization in tetrahydrofuran. Can J Chem 1967;45:1821-4.
[80] Yu YS, Dubois P, Teyssié P, Jérôme R. Difunctional initiator based on 1,3-
diisopropenylbenzene. 6. Synthesis of methyl methacrylate−butadiene−methyl
methacrylate triblock copolymers. Macromolecules 1997;30:4254-61.
[81] Yu JM, Dubois P, Jérôme R. Poly[methyl methacrylate (M)-b-styrene (S)-b-Butadiene
(B)-b-S-b-M] Pentablock Copolymers: Synthesis, Morphology, and Properties.
Macromolecules 1997;30:4984-94.
[82] Varshney S, Kesani P, Agarwal N, Zhang J, Rafailovich M. Synthesis of ABA type
thermoplastic elastomers based on polyacrylates. Macromolecules 1999;32:235-7.
[83] Tong J, Jérôme R. Synthesis of poly (methyl methacrylate)-b-poly (n-butyl acrylate)-b-
poly (methyl methacrylate) triblocks and their potential as thermoplastic elastomers.
Polymer 2000;41:2499-510.
[84] Moineau C, Minet M, Teyssié P, Jérôme R. Synthesis and characterization of
poly(methyl methacrylate)-block-poly(n-butyl acrylate)-block-poly(methyl
methacrylate) copolymers by two-step controlled radical polymerization (ATRP)
catalyzed by nibr2(pph3)2, 1. Macromolecules 1999;32:8277-82.
[85] Tong JD, Moineau G, Leclère Ph, Brédas JL, Lazzaroni R, Jérôme R. Synthesis,
morphology, and mechanical properties of poly(methyl methacrylate)-b-poly(n-butyl
acrylate)-b-poly(methyl methacrylate) triblocks. Ligated anionic polymerization vs atom
transfer radical polymerization. Macromolecules 2000;33:470-9.
77

[86] Tong JD, Jerôme R. Dependence of the ultimate tensile strength of thermoplastic
elastomers of the triblock type on the molecular weight between chain entanglements of
the central block. Macromolecules 2000;33:1479-81.
[87] Tong JD, Leclère P, Rasmont A, Brédas JL, Lazzaroni R, Jérôme R. Morphology and
rheology of poly (methyl methacrylate)-block-poly (isooctyl acrylate)-block-poly
(methyl methacrylate) triblock copolymers, and potential as thermoplastic elastomers.
Macromol Chem Phys 2000;201:1250-8.
[88] Leclère P, Rasmont A, Brédas JL, Jérôme R, Aime J, Lazzaroni R. Phase‐separated
microstructures in “all‐acrylic” thermoplastic elastomers. Macromol Symp
2001;167:117-37.
[89] Tong JD, Leclère P, Doneux C, Brédas JL, Lazzaroni R, Jérôme R. Morphology and
mechanical properties of poly(methylmethacrylate)-b-poly(alkylacrylate)-b-
poly(methylmethacrylate). Polymer 2001;42:3503-14.
[90] Chatterjee DP, Mandal BM. Triblock thermoplastic elastomers with poly (lauryl
methacrylate) as the center block and poly (methyl methacrylate) or poly (tert-butyl
methacrylate) as end blocks. Morphology and thermomechanical properties.
Macromolecules 2006;39:9192-200.
[91] Morishita Y. Applications of acrylic thermoplastic elastomer. Nippon Gomu Kyokaishi
2013;86:321-6.
[92] Baskaran D. Strategic developments in living anionic polymerization of alkyl
(meth)acrylates. Prog Polym Sci 2003;28:521-81.
[93] Ishizone T, Yoshimura K, Hirao A, Nakahama S. Controlled anionic polymerization of
tert-butyl acrylate with diphenylmethyl anions in the presence of dialkylzinc.
Macromolecules 1998;31:8706-12.
[94] Wang JS, Jerome R, Bayard P, Teyssie P. Anionic polymerization of acrylic monomers.
21. Anionic sequential polymerization of 2-ethylhexyl acrylate and methyl methacrylate.
Macromolecules 1994;27:4908-13.
[95] Leclère P, Moineau G, Minet M, Dubois P, Jérôme R, Brédas JL, et al. Direct
observation of microdomain morphology in “all-acrylic” thermoplastic elastomers
synthesized via living radical polymerization. Langmuir 1999;15:3915-9.
78

[96] Fetters LJ, Lohse DJ, Graessley WW. Chain dimensions and entanglement spacings in
dense macromolecular systems. J Polym Sci Part B Polym Phys 1999;37:1023-33.
[97] Bailey J, Bishop E, Hendrick W, Holden G, Legge N. Thermoplastic elastomers-
physical properties and applications. Rubber Age 1966;98:69-84.
[98] Luo Y, Wang X, Zhu Y, Li BG, Zhu S. Polystyrene-block-poly(n-butyl acrylate)-block-
polystyrene Triblock Copolymer Thermoplastic Elastomer Synthesized via RAFT
Emulsion Polymerization. Macromolecules 2010;43:7472-81.
[99] Hamada K, Morishita Y, Kurihara T, Ishiura K. Methacrylate-based polymers for
industrial uses. In: Hadjichristidis N, Hirao A, editors. Anionic Polymerization.
Heidelberg: Springer, 2015. p. 1011-31.
[100] Tabuchi M, Kawauchi T, Kitayama T, Hatada K. Living polymerization of primary
alkyl acrylates with t-butyllithium/bulky aluminum Lewis acids. Polymer 2002;43:7185-
90.
[101] Uchiumi N, Hamada K, Kato M, Ono T, Yaginuma S, Ishiura K. Preparation process of
acrylic acid ester polymer. US 6329480 B1, 2001.
[102] Hamada K, Ishiura K, Kato M, Yaginuma S. Anionic polymerization process, and
process for producing a polymer by the anionic polymerization process. EP 1078942 A1,
2006.
[103] Natori I. Synthesis of polymers with an alicyclic structure in the main chain. Living
anionic polymerization of 1, 3-cyclohexadiene with the n-butyllithium/N, N, N ‘, N ‘-
tetramethyl-ethylenediamine system. Macromolecules 1997;30:3696-7.
[104] Kosaka Y, Kitazawa K, Inomata S, Ishizone T. Living anionic polymerization of
benzofulvene: Highly reactive fixed transoid 1,3-diene. ACS Macro Lett 2013;2:164-7.
[105] Kosaka Y, Goseki R, Kawauchi S, Ishizone T. Living anionic polymerization of
benzofulvene in hydrocarbon solvent. Macromol Symp 2015;350:55-66.
[106] Imaizumi K, Ono T, Natori I, Sakurai S, Takeda K. Microphase‐separated structure of
1, 3‐cyclohexadiene/butadiene triblock copolymers and its effect on mechanical and
thermal properties. J Polym Sci Part B Polym Phys 2001;39:13-22.
[107] Natori I, Imaizumi K, Yamagishi H, Kazunori M. Hydrocarbon polymers containing
six‐membered rings in the main chain. Microstructure and properties of poly (1, 3‐
cyclohexadiene). J Polym Sci Part B Polym Phys 1998;36:1657-68.
79

[108] Kosaka Y, Kawauchi S, Goseki R, Ishizone T. High anionic polymerizability of


benzofulvene: New exo-methylene hydrocarbon monomer. Macromolecules
2015;48:4421-30.
[109] Wang W, Schlegel R, White BT, Williams K, Voyloy D, Steren CA, et al. High
temperature thermoplastic elastomers synthesized by living anionic polymerization in
hydrocarbon solvent at room temperature. Macromolecules 2016;49:2646-55.
[110] Legge NR. Thermoplastic elastomers. Rubber Chem Technol 1987;60:83-117.
[111] Hahn SF, Hillmyer MA. High glass transition temperature polyolefins obtained by the
catalytic hydrogenation of polyindene. Macromolecules 2003;36:71-6.
[112] Bates FS, Fredrickson GH, Hucul D, Hahn SF. PCHE-based pentablock copolymers:
Evolution of a new plastic. AIChE J 2001;47:762-5.
[113] Hucul DA, Hahn SF. Catalytic hydrogenation of polystyrene. Adv Mater 2000;12:1855-
8.
[114] Vigild ME, Chu C, Sugiyama M, Chaffin KA, Bates FS. Influence of shear on the
alignment of a lamellae-forming pentablock copolymer. Macromolecules 2001;34:951-
64.
[115] Alfonzo CG, Fleury G, Chaffin KA, Bates FS. Synthesis and characterization of
elastomeric heptablock terpolymers structured by crystallization. Macromolecules
2010;43:5295-305.
[116] Kaszas G, Puskas J, Kennedy J, Hager W. Polyisobutylene‐containing block polymers
by sequential monomer addition. II. Polystyrene–polyisobutylene–polystyrene
triblock polymers: Synthesis, characterization, and physical properties. J Polym Sci Part
A Polym Chem 1991;29:427-35.
[117] Cao X, Faust R. Polyisobutylene-based thermoplastic elastomers. 5. Poly (styrene-b-
isobutylene-b-styrene) triblock copolymers by coupling of living poly (styrene-b-
isobutylene) diblock copolymers. Macromolecules 1999;32:5487-94.
[118] Puskas JE, Kaszas G, Kennedy J, Hager W. Polyisobutylene‐containing block
polymers by sequential monomer addition. IV. New triblock thermoplastic elastomers
comprising high Tg styrenic glassy segments: Synthesis, characterization and physical
properties. J Polym Sci Part A Polym Chem 1992;30:41-8.
80

[119] Tsunogae Y, Kennedy J. Polyisobutylene‐containing block polymers by sequential


monomer addition. X. Synthesis of poly (α‐methylstyrene‐b‐isobutylene‐b‐α
‐methylstyrene) thermoplastic elastomers. J Polym Sci Part A Polym Chem
1994;32:403-12.
[120] Li D, Faust R. Polyisobutylene-based thermoplastic elastomers. 3. Synthesis,
characterization, and properties of poly (alpha.-methylstyrene-b-isobutylene-b-alpha-
methylstyrene) triblock copolymers. Macromolecules 1995;28:4893-8.
[121] Cao X, Sipos L, Faust R. Polyisobutylene based thermoplastic elastomers: VI. Poly (α-
methylstyrene-b-isobutylene-b-α-methylstyrene) triblock copolymers by coupling of
living poly (α-methylstyrene-b-isobutylene) diblock copolymers. Polym Bull
2000;45:121-8.
[122] Kennedy JP, Price JL, Koshimura K. Novel thermoplastic elastomer triblocks of a soft
polyisobutylene midblock connected to two hard PMMA stereocomplex outer blocks.
Macromolecules 1991;24:6567-71.
[123] Kennedy JP, Midha S, Tsunogae Y. Living carbocationic polymerization. 56.
Polyisobutylene-containing block polymers by sequential monomer addition. 8.
Synthesis, characterization, and physical properties of poly (indene-b-isobutylene-b-
indene) thermoplastic elastomers. Macromolecules 1993;26:429-35.
[124] Fodor Z, Kennedy J. Polyisobutylene-containing block polymers by sequential
monomer addition. Polym Bull 1992;29:697-704.
[125] Kennedy J, Kurian J. Living carbocationic polymerization of p‐halostyrenes. III.
Syntheses and characterization of novel thermoplastic elastomers of isobutylene and p‐
chlorostyrene. J Polym Sci Part A Polym Chem 1990;28:3725-38.
[126] Puskas JE, Gergely AL, Kaszas G. Controlled/living carbocationic copolymerization of
isobutylene with alloocimene. J Polym Sci Part A Polym Chem 2013;51:29-33.
[127] Gergely AL, Puskas JE. Synthesis and characterization of thermoplastic elastomers with
polyisobutylene and polyalloocimene blocks. J Polym Sci Part A Polym Chem
2015;53:1567-74.
[128] Roh JH, Roy D, Lee WK, Gergely AL, Puskas JE, Roland CM. Thermoplastic
elastomers of alloocimene and isobutylene triblock copolymers. Polymer 2015;56:280-3.
81

[129] Zhou Y, Faust R, Chen S, Gido SP. Synthesis, characterization, and morphology of poly
(tert-butyl vinyl ether-b-isobutylene-b-tert-butyl vinyl ether) triblock copolymers.
Macromolecules 2004;37:6716-25.
[130] Finch CA, editor. Polyvinyl Alcohol Developments. Chichester: John Wiley & Sons
Inc; 1992. 850 pp.
[131] Zhou Y, Faust R, Richard R, Schwarz M. Syntheses and characterization of poly
(cyclohexyl vinyl ether-s tat-vinyl alcohol)-b-polyisobutylene-b-poly (cyclohexyl vinyl
ether-s tat-vinyl alcohol) triblock copolymers and their application as coatings to deliver
paclitaxel from coronary stents. Macromolecules 2005;38:8183-91.
[132] Hashimoto T, Imaeda T, Irie S, Urushisaki M, Sakaguchi T. Synthesis of poly (vinyl
ether)‐based, ABA triblock‐type thermoplastic elastomers with functionalized soft
segments and their gas permeability. J Polym Sci Part A Polym Chem 2015;53:1114-24.
[133] Imaeda T, Hashimoto T, Irie S, Urushisaki M, Sakaguchi T. Synthesis of ABA‐
triblock and star‐diblock copolymers with poly (2‐adamantyl vinyl ether) and poly
(n‐butyl vinyl ether) segments: New thermoplastic elastomers composed solely of poly
(vinyl ether) backbones. J Polym Sci Part A Polym Chem 2013;51:1796-807.
[134] Sakaguchi T, Okunaga R, Irie S, Urushisaki M, Hashimoto T. Carbon dioxide-
permselective polymer membranes composed of poly(vinyl ether)-based, ABA-type
triblock copolymers with pendant oxyethylene chains. Polym Bull 2017;74:2017-31.
[135] Robeson LM. The upper bound revisited. J Membr Sci 2008;320:390-400.
[136] Powell CE, Qiao GG. Polymeric CO2/N2 gas separation membranes for the capture of
carbon dioxide from power plant flue gases. J Membr Sci 2006;279:1-49.
[137] Pan P, Inoue Y. Polymorphism and isomorphism in biodegradable polyesters. Prog
Polym Sci 2009;34:605-40.
[138] Wang Z, Yuan L, Tang C. Sustainable elastomers from renewable biomass. Acc Chem
Res 2017;50:1762-73.
[139] Hillmyer MA, Tolman WB. Aliphatic polyester block polymers: Renewable,
degradable, and sustainable. Acc Chem Res 2014;47:2390-6.
[140] Belgacem MN, Gandini A, editors. Monomers, polymers and composites from
renewable resources. Oxford: Elsevier; 2011. 560 pp.
82

[141] Frick EM, Zalusky AS, Hillmyer MA. Characterization of polylactide-b-polyisoprene-


b-polylactide thermoplastic elastomers. Biomacromolecules 2003;4:216-23.
[142] Sipos L, Zsuga M, Deák G. Synthesis of poly (L‐lactide)‐block‐polyisobutylene‐
block‐poly (L‐lactide), a new biodegradable thermoplastic elastomer. Macromol
Rapid Comm 1995;16:935-40.
[143] Zhang S, Hou Z, Gonsalves K. Copolymer synthesis of poly (L-lactide-b-DMS-L-
lactide) via the ring opening polymerization of L-lactide in the presence of alpha,
omega-hydroxylpropyl-terminated PDMS macroinitiator. J Polym Sci Part A Polym
Chem 1996;34:2737-42.
[144] Han L, Yu C, Zhou J, Shan G, Bao Y, Yun X, et al. Enantiomeric blends of high-
molecular-weight poly(lactic acid)/poly(ethylene glycol) triblock copolymers: Enhanced
stereocomplexation and thermomechanical properties. Polymer 2016;103:376-86.
[145] Huang Y, Pan P, Shan G, Bao Y. Polylactide-b-poly (ethylene-co-butylene)-b-
polylactide thermoplastic elastomers: role of polylactide crystallization and
stereocomplexation on microphase separation, mechanical and shape memory
properties. RSC Adv 2014;4:47965-76.
[146] Tsuji H. Poly(lactide) stereocomplexes: Formation, structure, properties, degradation,
and applications. Macromol Biosci 2005;5:569-97.
[147] Wanamaker CL, Tolman WB, Hillmyer MA. Poly(D-lactide)–poly(menthide)–poly(D-
lactide) triblock copolymers as crystal nucleating agents for poly(L-lactide). Macromol
Symp 2009;283-284:130-8.
[148] Wanamaker CL, O’Leary LE, Lynd NA, Hillmyer MA, Tolman WB. Renewable-
resource thermoplastic elastomers based on polylactide and polymenthide.
Biomacromolecules 2007;8:3634-40.
[149] Wanamaker CL, Bluemle MJ, Pitet LM, O’Leary LE, Tolman WB, Hillmyer MA.
Consequences of polylactide stereochemistry on the properties of polylactide-
polymenthide-polylactide thermoplastic elastomers. Biomacromolecules 2009;10:2904-
11.
[150] Xiong M, Schneiderman DK, Bates FS, Hillmyer MA, Zhang K. Scalable production of
mechanically tunable block polymers from sugar. Proc Nat Acad Sci 2014;111:8357-62.
83

[151] Watts A, Kurokawa N, Hillmyer MA. Strong, resilient, and sustainable aliphatic
polyester thermoplastic elastomers. Biomacromolecules 2017;18:1845-54.
[152] Schneiderman DK, Hillmyer MA. Aliphatic polyester block polymer design.
Macromolecules 2016;49:2419-28.
[153] Schneiderman DK, Hill EM, Martello MT, Hillmyer MA. Poly (lactide)-block-poly (ε-
caprolactone-co-ε-decalactone)-block-poly (lactide) copolymer elastomers. Polym Chem
2015;6:3641-51.
[154] Nakayama Y, Aihara K, Yamanishi H, Fukuoka H, Tanaka R, Cai Z, et al. Synthesis of
biodegradable thermoplastic elastomers from ε‐caprolactone and lactide. J Polym Sci
Part A Polym Chem 2015;53:489-95.
[155] Martello MT, Hillmyer MA. Polylactide–poly (6-methyl-ε-caprolactone)–polylactide
thermoplastic elastomers. Macromolecules 2011;44:8537-45.
[156] Matyjaszewski K, Davis TP. Handbook of Radical Polymerization. New York:John
Wiley & Sons, 2003. 920 pp.
[157] Matyjaszewski K, Tsarevsky NV. Nanostructured functional materials prepared by
atom transfer radical polymerization. Nat Chem 2009;1:276-88.
[158] Matyjaszewski K. Architecturally complex polymers with controlled heterogeneity.
Science 2011;333:1104-5.
[159] Dong H, Tang W, Matyjaszewski K. Well-defined high-molecular-weight
polyacrylonitrile via activators regenerated by electron transfer ATRP. Macromolecules
2007;40:2974-7.
[160] Thomas D, Sumerlin B, Lowe A, McCormick C. Conditions for facile, controlled
RAFT polymerization of acrylamide in water. Macromolecules 2003;36:1436-9.
[161] Keddie DA. Guide to the synthesis of block copolymers using reversible-addition
fragmentation chain transfer (RAFT) polymerization. Chem Soc Rev 2014;43:496-505.
[162] Suenaga J, Sutherlin DM, Stille JK. Polymerization of (RS)-and (R)-α-methylene-γ-
methyl-γ-butyrolactone. Macromolecules 1984;17:2913-6.
[163] Akkapeddi MK. The free radical copolymerization characteristics of α-methylene γ-
butyrolactone. Polymer 1979;20:1215-6.
[164] Shin J, Lee Y, Tolman WB, Hillmyer MA. Thermoplastic elastomers derived from
menthide and tulipalin A. Biomacromolecules 2012;13:3833-40.
84

[165] Satoh K, Lee DH, Nagai K, Kamigaito M. Precision Synthesis of Bio-Based Acrylic
Thermoplastic Elastomer by RAFT Polymerization of Itaconic Acid Derivatives.
Macromol Rapid Commun 2014;35;161-67.
[166] Haloi DJ, Ata S, Singha NK, Jehnichen D, Voit B. Acrylic AB and ABA block
copolymers based on poly(2-ethylhexyl acrylate) (PEHA) and poly(methyl
methacrylate) (PMMA) via ATRP. ACS Appl Mater Interfaces 2012;4:4200-7.
[167] Robin S, Guerret O, Couturier JL, Pirri R, Gnanou Y. Synthesis and characterization of
poly(styrene-b-n-butyl acrylate-b-styrene) triblock copolymers using a dialkoxyamine as
initiator. Macromolecules 2002;35:3844-8.
[168] Nicolas J, Charleux B, Guerret O, Magnet S. Nitroxide-mediated controlled free-radical
emulsion polymerization using a difunctional water-soluble alkoxyamine initiator.
Toward the control of particle size, particle size distribution, and the synthesis of
triblock copolymers. Macromolecules 2005;38:9963-73.
[169] Wang S, Kesava SV, Gomez ED, Robertson ML. Sustainable thermoplastic elastomers
derived from fatty acids. Macromolecules 2013;46:7202-12.
[170] Kang NG, Lu W, Wang Y, Wang W, Cheng S, Zhu J, et al. All acrylic-based
thermoplastic elastomers with high upper service temperature and superior mechanical
properties. Polym Chem 2017;8:5741-8
[171] Zetterlund PB, Thickett SC, Perrier Sb, Bourgeat-Lami E, Lansalot M, Thickett SC, et
al. Controlled/living radical polymerization in dispersed systems: An update. Chem Rev
2015;115:9745-800.
[172] Zhan X, He R, Zhang Q, Chen F. Microstructure and mechanical properties of
amphiphilic tetrablock copolymer elastomers via RAFT miniemulsion polymerization:
in fl uence of poly[styrene- alt -(maleic anhydride)] segments. RSC Adv 2014;4:51201-
7.
[173] Wei R, Luo Y, Zeng W, Wang F, Xu S. Styrene-butadiene−styrene triblock copolymer
latex via reversible addition−fragmentation chain transfer miniemulsion polymerization.
Ind Eng Chem Res 2012;51:15530-5.
[174] Vasile C. Handbook of polyolefins. 2nd Ed. New York: CRC Press; 2000. 1032 pp.
[175] Natta G. Properties of isotactic, atactic, and stereoblock homopolymers, random and
block copolymers of α‐olefins. J Polym Sci 1959;34:531-49.
85

[176] Collette JW, Tullock CW, MacDonald RN, Buck WH, Su ACL, Harrell JR, et al.
Elastomeric polypropylenes from alumina-supported tetraalkyl group IVB catalysts. 1.
Synthesis and properties of high molecular weight stereoblock homopolymers.
Macromolecules 1989;22:3851-8.
[177] Collette JW, Ovenall DW, Buck WH, Fergusont RC. Elastomeric polypropylenes from
alumina-supported tetraalkyl group IVB catalysts. 2. Chain microstructure, crystallinity,
and morphology. Macromolecules 1989;22:3858-66
[178] Mallín DT, Rausch MD, Lin YG, Dong S, Chien JCW. rae-[Ethy lidene( 1 -r)5-
tetramethylcyclopentadienyl) (1- 5-indenyl)]dichlorotitanium and its
homopolymerization of propylene to crystalline-amorphous block thermoplastic
elastomers. J Am Chem Soc 1990;112:2030-1.
[179] Dietrich U, Hackmann M, Rieger B, Klinga M, Leskelä M. Control of stereoerror
formation with high-activity “dual-side” zirconocene catalysts:  A novel strategy to
design the properties of thermoplastic elastic polypropenes. J Am Chem Soc
1999;121:4348-55.
[180] Coates GW, Waymouth RM. Oscillating stereocontrol: a strategy for the synthesis of
thermoplastic elastomeric polypropylene. Science 1995;267:217-9.
[181] Machat MR, Lanzinger D, Drees M, Altmann PJ, Herdtweck E, Rieger B. High-
Melting, Elastic Polypropylene: A One-Pot, One-Catalyst Strategy toward Propylene-
Based Thermoplastic Elastomers. Macromolecules 2018;51:914-29.
[182] Moshonov M, Aharonovich S, Eisen MS. Tailor-Made Thermoplastic Elastomeric
Stereoblock Polypropylenes by Modulation of Monomer Pressure. Macromolecules
2016;49:9287-90.
[183] Zhang Y, Keaton RJ, Sita LR. Degenerative transfer living ziegler - natta
polymerization: Application to the synthesis of monomodal stereoblock polyolefins of
narrow polydispersity and tunable block length. J Am Chem Soc 2003;125:9062-9.
[184] Zhang Y, Sita LR. Stereospecific Living Ziegler - Natta Polymerization via Rapid and
Reversible Chloride Degenerative Transfer between Active and Dormant Sites. J Am
Chem Soc 2004;126:7776 - 7.
[185] Harney MB, Zhang Y, Sita LR. Discrete, multiblock isotactic–atactic stereoblock
polypropene microstructures of differing block architectures through programmable
86

stereomodulated living Ziegler–Natta polymerization. Angew Chem Int Ed


2006;45:2400-4.
[186] Hotta A, Cochran E, Ruokolainen J, Khanna V, Fredrickson GH, Kramer EJ, et al.
Semicrystalline thermoplastic elastomeric polyolefins: Advances through catalyst
development and macromolecular design. Proc Nat Acad Sci 2006;103:15327-32.
[187] Crawford KE, Sita LR. Stereoengineering of poly(1,3-methylenecyclohexane) via two-
state living coordination polymerization of 1,6-heptadiene. J Am Chem Soc
2013;135:8778-81.
[188] Crawford KE, Sita LR. De Novo design of a new class of “hard-soft” amorphous,
microphase-separated, polyolefin block copolymer thermoplastic elastomers. ACS
Macro Lett 2015;4:921-5.
[189] Guan Z, Cotts PM, McCord EF, McLain SJ. Chain walking: a new strategy to control
polymer topology. Science 1999;283:2059-62.
[190] Leone G, Mauri M, Pierro I, Ricci G, Canetti M, Bertini F. Polyolefin thermoplastic
elastomers from 1-octene chain-walking polymerization. Polymer 2016;100:37-44.
[191] Lian K, Zhu Y, Li W, Dai S, Chen C. Direct synthesis of thermoplastic polyolefin
elastomers from nickel-catalyzed ethylene polymerization. Macromolecules
2017;50:6074-80.
[192] O’Connor KS, Watts A, Vaidya T, LaPointe AM, Hillmyer MA, Coates GW.
Controlled chain walking for the synthesis of thermoplastic polyolefin elastomers:
synthesis, structure, and properties. Macromolecules 2016;49:6743-51.
[193] Watanabe H. Slow dielectric relaxation of a styrene-isoprene-styrene triblock
copolymer with dipole inversion in the middle block: A challenge to a loop/bridge
problem. Macromolecules 1995;28:5006-11.
[194] Takahashi Y, Song Y, Nemoto N, Takano A, Akazawa Y, Matsushita Y. Effect of
loop/bridge conformation ratio on elastic properties of the sphere-forming aba triblock
copolymers under uniaxial elongation. Macromolecules 2005;38:9724-9.
[195] Takano A, Kamaya I, Takahashi Y, Matsushita Y. Effect of loop/bridge conformation
ratio on elastic properties of the sphere-forming aba triblock copolymers: Preparation of
samples and determination of loop/bridge ratio. Macromolecules 2005;38:9718-23.
87

[196] Matsen MW, Schick M. Lamellar phase of a symmetric triblock copolymer.


Macromolecules 1994;27:187-92
[197] Matsen MW, Thompson RB. Equilibrium behavior of symmetric ABA triblock
copolymer melts. J Chem Phys 1999;111:7139-46.
[198] Brinkmann-Rengel S, Abetz V, Stadler R, Thomas E. Thermoplastic elastomers based
on ABA-and ABC-triblock copolymers. Kautsch Gummi Kunstst 1999;52:806-13.
[199] Stadler R, Auschra C, Beckmann J, Krappe U, Voight-Martin I, Leibler L. Morphology
and thermodynamics of symmetric poly (A-block-B-block-C) triblock copolymers.
Macromolecules 1995;28:3080-97.
[200] Auschra C, Stadler R. New ordered morphologies in ABC triblock copolymers.
Macromolecules 1993;26:2171-4.
[201] Abetz V, Stadler R, Leibler L. Order-disorder-and order-order-transitions in AB and
ABC block copolymers: description by a simple model. Polym Bull 1996;37:135-42.
[202] Kabir MR. Correlation between morphology and mechanical properties of polystyrene-
b-polybutadiene-b-poly(methacrylate) in dependency of the polybutadiene’s block
microstructure. RNDr Thesis. Geesthacht, Germany: Helmholtz-Zentrum Geesthacht
Centre for Materials and Coastal Research, 2011. 241 pp.
[203] Kabir R, Albuerne J, Simon PFW, Filiz V, Abetz C, Böttcher H, et al. Deformation and
orientation behavior of polystyrene-b-polybutadiene-b-poly(methyl methacrylate)
triblock terpolymers: Influence of polybutadiene microstructures and the molar masses.
Polymer 2013;54:673-84.
[204] Rangarajan P, Register RA, Fetters LJ. Morphology of semicrystalline block
copolymers of ethylene-(ethylene-alt-propylene). Macromolecules 1993;26:4640-5.
[205] Kofinas P, Cohen RE. Morphology of highly textured poly (ethylene)/poly (ethylene-
propylene)(E/EP) semicrystalline diblock copolymers. Macromolecules 1994;27:3002-8.
[206] Rangarajan P, Register RA, Adamson DH, Fetters LJ, Bras W, Naylor S, et al.
Dynamics of structure formation in crystallizable block copolymers. Macromolecules
1995;28:1422-8.
[207] Balsamo V, Navarro CUd, Gil G. Microphase separation vs crystallization in
polystyrene-b-polybutadiene-b-poly( -caprolactone) abc triblock copolymers.
Macromolecules 2003 :4507-14.
88

[208] Balsamo V, von Gyldenfeldt F, Stadler R. Synthesis of SBC, SC and BC block


copolymers based on polystyrene (S), polybutadiene (B) and a crystallizable poly (ε‐
caprolactone)(C) block. Macromol Chem Phys 1996;197:1159-69.
[209] Balsamo V, von Gyldenfeldt F, Stadler R. “Superductile” semicrystalline abc triblock
copolymers with the polystyrene block (a) as the matrix. Macromolecules
1999;32:1226-32.
[210] Balsamo V, Gil G, Navarro CUd, Hamley IW, Gyldenfeldt Fv, Abetz V, et al.
Morphological behavior of thermally treated polystyrene-b-polybutadiene-b-poly( -
caprolactone) abc triblock copolymers. Macromolecules 2003;36:4515-25.
[211] Schmalz H, Böker A, Lange R, Krausch G, Abetz V. Synthesis and properties of ABA
and ABC triblock copolymers with glassy (A), elastomeric (B), and crystalline (C)
blocks. Macromolecules 2001;34:8720-9.
[212] Schmalz H, Abetz V, Lange R. Thermoplastic elastomers based on semicrystalline
block copolymers. Compos Sci Technol 2003;63:1179-86.
[213] Aoyagi T, Honda T, Doi M. Microstructural study of mechanical properties of the ABA
triblock copolymer using self-consistent field and molecular dynamics. J Chem Phys
2002;117:8153-61.
[214] Hermel TJ, Hahn SF, Chaffin KA, Gerberich WW, Bates FS. Role of molecular
architecture in mechanical failure of glassy/semicrystalline block copolymers: CEC vs
CECEC lamellae. Macromolecules 2003;36:2190-3.
[215] Mori Y, Lim LS, Bates FS. Consequences of molecular bridging in lamellae-forming
triblock/pentablock copolymer blends. Macromolecules 2003;36:9879-88.
[216] Wu L, Cochran EW, Lodge TP, Bates FS. Consequences of block number on the order -
disorder transition and viscoelastic properties of linear (AB)n multiblock copolymers.
Macromolecules 2004;37:3360-8.
[217] Koo CM, Hillmyer MA, Bates FS. Structure and properties of semicrystalline - rubbery
multiblock copolymers. Macromolecules 2006;39:667-77.
[218] Fleury G, Bates FS. Structure and properties of hexa- and undecablock terpolymers
with hierarchical molecular architectures. Macromolecules 2009;42:3598 - 610.
[219] Lee I, Bates FS. Synthesis, structure, and properties of alternating and random

poly(styrene‑b‑butadiene) multiblock copolymers. Macromolecules 2013;46:4529-39.


89

[220] Lee I, Panthani TR, Bates FS. Sustainable poly(lactide‑b‑butadiene) multiblock

copolymers with enhanced mechanical properties. Macromolecules 2013;46:7387-98.


[221] Panthani TR, Bates FS. Crystallization and mechanical properties of poly(L-lactide)-
based rubbery/semicrystalline multiblock copolymers. Macromolecules 2015;48:4529-
40.
[222] Walker CN, Sarapas JM, Kung V, Hall AL, Tew GN. Multiblock copolymers by thiol
addition across norbornene. ACS Macro Lett 2014;3:453-7.
[223] Zhang D, Dumont MJ. Synthesis, characterization and potential applications of 5-
hydroxymethylfurfural derivative based poly( β-thioether esters) synthesized via thiol-
Michael addition polymerization. Polym Chem 2018;9:743-56.
[224] Bell CA, Yu J, Barker IA, Truong VX, Cao Z, Dobrinyin AV, et al. Independent control
of elastomer properties through stereocontrolled synthesis. Angew Chem Int Ed
2016;55:13076-80.
[225] Li H, Thanneeru S, Jin L, Guild CJ, He J. Multiblock thermoplastic elastomers via one-
pot thiol-ene reaction. Polym Chem 2016;7:4824-32.
[226] Moon NG, Mondschein RJ, Long TE. Poly( β -thioesters) containing monodisperse
oxamide hard segments using a chemoselective thiol-michael addition reaction. Polym
Chem 2017;8:2598-608.
[227] Jin K, Leitsch EK, Chen X, Heath WH, Torkelson JM. Segmented thermoplastic
polymers synthesized by thiol−ene click chemistry: Examples of thiol−norbornene and
thiol−maleimide click reactions. Macromolecules 2018;51:3620-31.
[228] Arriola DJ, Carnahan EM, Hustad PD, Kuhlman RL, Wenzel TT. Catalytic production
of olefin block copolymers via chain shuttling polymerization. Science 2006;312:714-9.
[229] Chum PS, Swogger KW. Olefin polymer technologies—history and recent progress at
the Dow Chemical Company. Prog Polym Sci 2008;33:797-819.
[230] Jandaghian MH, Soleimannezhad A, Ahmadjo S, Mortazavi SMM, Ahmadi M.
Synthesis and characterization of isotactic poly(1-hexene)/branched polyethylene
multiblock copolymer via chain shuttling polymerization technique. Ind Eng Chem Res
2018;57:4807-14.
[231] Valente A, Stoclet G, Bonnet F, Mortreux A, Visseaux M, Zinck P. Isoprene-styrene
chain shuttling copolymerization mediated by a lanthanide half-sandwich complex and a
90

lanthanidocene: Straightforward access to a new type of thermoplastic elastomers.


Angew Chem Int Ed 2014;53:4638 -41.
[232] You L, Ling J. Janus polymerization. Macromolecules 2014;47:2219-25.
[233] Ling J, Wang X, You L, Shen Z. thermoplastic elastomers based on poly(L-lysine)-
poly(ε-caprolactone) multi-block copolymers. J Polym Sci Part A Polym Chem
2016;54:3012-8.
[234] Hu N, Mai CK, Fredrickson GH, Bazan GC. One-pot synthesis of
semicrystalline/amorphous multiblock copolymers via divinyl-terminated telechelic
polyolefins. Chem Commun 2016;52:2237-40.
[235] Kim DH, Park SS, Park SH, Jeon JY, Kim HB, Lee BY. Preparation of polystyrene -
polyolefin multiblock copolymers by sequential coordination and anionic
polymerization. RSC Adv 2017;7:5948-56.
[236] Hadjichristidis N, Pitsikalis M, Pispas S, Iatrou H. Polymers with complex architecture
by living anionic polymerization. Chem Rev 2001;101:3747-92.
[237] Khanna K, Varshney S, Kakkar A. Miktoarm star polymers: advances in synthesis, self-
assembly, and applications. Polym Bull 2010;1:1171-85.
[238] Hadjichristidis N. Synthesis of miktoarm star (μ‐star) polymers. J Polym Sci Part A
Polym Chem 1999;37:857-71.
[239] Bi LK, Fetters LJ. Synthesis and properties of block copolymers. 3. Polystyrene-
polydiene star block copolymers. Macromolecules 1976;9:732-42.
[240] Bi LK, Fetters L. Domain morphology of star block copolymers of polystyrene and
polyisoprene. Macromolecules 1975;8:90-2.
[241] Spencer RK, Matsen MW. Domain Bridging in Thermoplastic Elastomers of Star Block
Copolymer. Macromolecules 2017;50:1681-7.
[242] Dair BJ, Honeker CC, Alward DB, Avgeropoulos A, Hadjichristidis N, Fetters LJ, et al.
Mechanical properties and deformation behavior of the double gyroid phase in
unoriented thermoplastic elastomers. Macromolecules 1999;32:8145-52.
[243] Lai C, Russel WB, Register RA, Marchand GR, Adamson DH. Phase behavior of
styrene-isoprene diblock derivatives with varying conformational asymmetry.
Macromolecules 2000;33:3461-6.
91

[244] Milner ST. Chain architecture and asymmetry in copolymer microphases.


Macromolecules 1994;27:2333-5.
[245] Vavasour J, Whitmore M. Self-consistent field theory of block copolymers with
conformational asymmetry. Macromolecules 1993;26:7070-5.
[246] Lynd NA, Oyerokun FT, O’Donoghue DL, Handlin Jr DL, Fredrickson GH. Design of
soft and strong thermoplastic elastomers based on nonlinear block copolymer
architectures using self-consistent-field theory. Macromolecules 2010;43:3479-86.
[247] Bates FS, Fredrickson GH. Block copolymer thermodynamics: theory and experiment.
Annu Rev Phys Chem 1990;41:525-57.
[248] Shi W, Lynd NA, Montarnal D, Luo Y, Fredrickson GH, Kramer EJ, et al. Toward
strong thermoplastic elastomers with asymmetric miktoarm block copolymer
architectures. Macromolecules 2014;47:2037-43.
[249] Michler GH, Adhikari R, Lebek W, Goerlitz S, Weidisch R, Knoll K. Morphology and
micromechanical deformation behavior of styrene/butadiene-block copolymers. I.
Toughening mechanisms in asymmetric star block copolymers. J Appl Polym Sci
2002;85:683-700.
[250] Shi W, Hamilton AL, Delaney KT, Fredrickson GH, Kramer EJ, Ntaras C, et al.
Aperiodic “bricks and mortar” mesophase: A new equilibrium state of soft matter and
application as a stiff thermoplastic elastomer. Macromolecules 2015;48:5378-84.
[251] Shi W, Hamilton AL, Delaney KT, Fredrickson GH, Kramer EJ, Ntaras C, et al.
Creating extremely asymmetric lamellar structures via fluctuation-assisted unbinding of
miktoarm star block copolymer alloys. J Am Chem Soc 2015;137:6160-3.
[252] Matmour R, Gnanou Y. Synthesis of complex polymeric architectures using
multilithiated carbanionic initiators—Comparison with other approaches. Prog Polym
Sci 2013;38:30-62.
[253] Theodosopoulos GV, Hurley CM, Mays JW, Sakellariou G, Baskaran D. Trifunctional
organolithium initiator for living anionic polymerization in hydrocarbon solvents in the
absence of polar additives. Polym Chem 2016;7:4090-9.
[254] Jacob S, Kennedy JP. Synthesis and characterization of novel octa-arm star-block
thermoplastic elastomers consisting of poly (p-chlorostyrene-b-isobutylene) arms
radiating from a calix[8]arene core. Polym Bull 1998;41:167-74.
92

[255] Keszler B, Fenyvesi G, Kennedy J. Novel star‐block polymers: Three


polyisobutylene‐b‐poly (methyl methacrylate) arms radiating from an aromatic core.
J Polym Sci Part A Polym Chem 2000;38:706-14.
[256] Shim JS, Asthana S, Omura N, Kennedy JP. Novel thermoplastic elastomers. I.
Synthesis and characterization of star‐block copolymers of PSt‐b‐PIB arms
emanating from cyclosiloxane cores. J Polym Sci Part A Polym Chem 1998;36:2997-
3012.
[257] Shim JS, Kennedy JP. Novel thermoplastic elastomers. II. Properties of star‐block
copolymers of PSt‐b‐PIB arms emanating from cyclosiloxane cores. J Polym Sci Part
A Polym Chem 1999;37:815-24.
[258] Shim JS, Kennedy JP. Novel thermoplastic elastomers. III. Synthesis, characterization,
and properties of star‐block copolymers of poly (indene‐b‐isobutylene) arms
emanating from cyclosiloxane cores. J Polym Sci Part A Polym Chem 2000;38:279-90.
[259] Juhari A, Mosnáček J, Yoon JA, Nese A, Koynov K, Kowalewski T, et al. Star-like
poly (n-butyl acrylate)-b-poly (α-methylene-γ-butyrolactone) block copolymers for high
temperature thermoplastic elastomers applications. Polymer 2010;51:4806-13.
[260] Dufour B, Koynov K, Pakula T, Matyjaszewski K. PBA–PMMA 3‐arm star block
copolymer thermoplastic elastomers. Macromol Chem Phys 2008;209:1686-93.
[261] Dufour B, Tang C, Koynov K, Zhang Y, Pakula T, Matyjaszewski K. Polar three-arm
star block copolymer thermoplastic elastomers based on polyacrylonitrile.
Macromolecules 2008;41:2451-8.
[262] Pakula T, Koynov K, Boerner H, Huang J, Lee HI, Pietrasik J, et al. Effect of chain
topology on the self-organization and the mechanical properties of poly (n-butyl
acrylate)-b-polystyrene block copolymers. Polymer 2011;52:2576-83.
[263] Hadjichristidis N, Pispas S, Floudas G. Block Copolymers: Synthetic Strategies,
Physical Properties, And Applications. New Jersey: John Wiley & Sons; 2003. 440 pp.
[264] Wang H, Lu W, Wang W, Shah PN, Misichronis K, Kang NG, et al. Design and
synthesis of multigraft copolymer thermoplastic elastomers: Superelastomers. Macromol
Chem Phys 2018;219:1700254/1-11.
93

[265] Iatrou H, Mays JW, Hadjichristidis N. Regular comb polystyrenes and graft
polyisoprene/polystyrene copolymers with double branches (“centipedes”). Quality of
(1, 3-phenylene) bis (3-methyl-1-phenylpentylidene) dilithium initiator in the presence
of polar additives. Macromolecules 1998;31:6697-701.
[266] Mays JW, Uhrig D, Gido S, Zhu Y, Weidisch R, Iatrou H, et al. Synthesis and
structure–property relationships for regular multigraft copolymers. Macromol Symp
2004;215:111-26.
[267] Uhrig D, Mays J. Synthesis of well-defined multigraft copolymers. Polym Chem
2011;2:69-76.
[268] Uhrig D, Mays JW. Synthesis of combs, centipedes, and barbwires: poly (isoprene-
graft-styrene) regular multigraft copolymers with trifunctional, tetrafunctional, and
hexafunctional branch points. Macromolecules 2002;35:7182-90.
[269] Mijović J, Sun M, Pejanović S, Mays JW. Effect of molecular architecture on dynamics
of multigraft copolymers: combs, centipedes, and barbwires. Macromolecules
2003;36:7640-51.
[270] Zhu Y, Burgaz E, Gido SP, Staudinger U, Weidisch R, Uhrig D, et al. Morphology and
tensile properties of multigraft copolymers with regularly spaced tri-, tetra-, and
hexafunctional junction points. Macromolecules 2006;39:4428-36.
[271] Duan Y, Rettler E, Schneider K, Schlegel R, Thunga M, Weidisch R, et al. Deformation
behavior of sphere-forming trifunctional multigraft copolymer. Macromolecules
2008;41:4565-8.
[272] Duan Y, Thunga M, Schlegel R, Schneider K, Rettler E, Weidisch R, et al. Morphology
and deformation mechanisms and tensile properties of tetrafunctional multigraft
copolymers. Macromolecules 2009;42:4155-64.
[273] Schlegel R, Wilkin D, Duan Y, Weidisch R, Heinrich G, Uhrig D, et al. Stress softening
of multigraft copolymers. Polymer 2009;50:6297-304.
[274] Staudinger U, Weidisch R, Zhu Y, Gido S, Uhrig D, Mays J, et al. Mechanical
properties and hysteresis behaviour of multigraft copolymers. Macromol Symp
2006;233: 42-50.
[275] Uhrig D, Schlegel R, Weidisch R, Mays J. Multigraft copolymer superelastomers:
Synthesis morphology, and properties. Eur Polym J 2011;47:560-8.
94

[276] Wang W, Wang W, Lu X, Bobade S, Chen J, Kang NG, et al. Synthesis and
characterization of comb and centipede multigraft copolymers pnba-g-ps with high
molecular weight using miniemulsion polymerization. Macromolecules 2014;47:7284-
95.
[277] Goodwin A, Wang W, Kang NG, Wang Y, Hong K, Mays J. All-acrylic multigraft
copolymers: Effect of side chain molecular weight and volume fraction on mechanical
behavior. Ind Eng Chem Res 2015;54:9566-76.
[278] Li H, Wang W, Li C, Tan J, Yin D, Zhang H, et al. Synthesis and characterization of
brush-like multigraft copolymers PnBA-g-PMMA by a combination of emulsion AGET
ATRP and emulsion polymerization. J Colloid Interface Sci 2015;453:226-36.
[279] Li H, Wang W, Tan J, Li C, Zhang Q. Synthesis and characterization of graft
copolymers PnBA-g-PS by miniemulsion polymerization. RSC Adv 2015;5:45459-66.
[280] Wang W, Wang W, Li H, Lu X, Chen J, Kang NG, et al. Synthesis and characterization
of graft copolymers poly(isoprene-g-styrene) of high molecular weight by a combination
of anionic polymerization and emulsion polymerization. Ind Eng Chem Res
2015;54:1292-300.
[281] Lu W, Goodwin A, Wang Y, Yin P, Wang W, Zhu J, et al. All-acrylic superelastomers:
facile synthesis and exceptional mechanical behavior. Polym Chem 2018;9:160-8.
[282] Fónagy T, Iván B, Szesztay M. Polyisobutylene-graft-polystyrene by quasiliving atom
transfer radical polymerization of styrene from poly(isobutylene-co-p-methylstyrene-co-
p-bromomethylstyrene). Macromol Rapid Comm 1998;19:479-83.
[283] Suksawad P, Yamamoto Y, Kawahara S. Preparation of thermoplastic elastomer from
natural rubber grafted with polystyrene. Eur Polym J 2011;47:330-7.
[284] Zhou C, Wei Z, Wang Y, Yu Y, Leng X, Li Y. Fully biobased thermoplastic
elastomers: Synthesis of highly branched star comb poly (β-myrcene)-graft-poly (l-
lactide) copolymers with tunable mechanical properties. Eur Polym J 2018;99:477-84.
[285] Lichtenhan JD. Polyhedral oligomeric silsesquioxanes: building blocks for
silsesquioxane-based polymers and hybrid materials. Comments Inorg Chem
1995;17:115-30.
95

[286] Zheng L, Hong S, Cardoen G, Burgaz E, Gido SP, Coughlin EB. Polymer
nanocomposites through controlled self-assembly of cubic silsesquioxane scaffolds.
Macromolecules 2004;37:8606-11.
[287] Seurer B, Coughlin EB. Ethylene–propylene–silsesquioxane thermoplastic elastomers.
Macromol Chem Phys 2008;209:1198-209.
[288] Schneider Y, Lynd NA, Kramer EJ, Bazan GC. Novel elastomers prepared by grafting
n-butyl acrylate from polyethylene macroinitiator copolymers. Macromolecules
2009;42:8763-8.
[289] Coffin RC, Schneider Y, Kramer EJ, Bazan GC. Binuclear initiators for the telechelic
synthesis of elastomeric polyolefins. J Am Chem Soc 2010;132:13869-78.
[290] Jiang F, Wang Z, Qiao Y, Wang Z, Tang C. A novel architecture toward third-
generation thermoplastic elastomers by a grafting strategy. Macromolecules
2013;46:4772-80.
[291] Lu C, Yu J, Wang C, Wang J, Chu F. Fabrication of UV-absorbent cellulose-rosin
based thermoplastic elastomer via “graft from” ATRP. Carbohydr Polym 2018;188:128-
35.
[292] Liu Y, Yao K, Chen X, Wang J, Wang Z, Ploehn HJ, Wang C, Chu F, Tang C.
Sustainable thermoplastic elastomers derived from renewable cellulose, rosin and fatty
acids. Polym Chem 2014;5:3170-81.
[293] Wang Z, Jiang F, Zhang Y, You Y, Wang Z, Guan Z. Bioinspired design of
nanostructured elastomers with cross-linked soft matrix grafting on the oriented rigid
nanofibers to mimic mechanical properties of human skin. ACS Nano 2014;9:271-8.
[294] Wang Z, Zhang Y, Jiang F, Fang H, Wang Z. Synthesis and characterization of
designed cellulose-graft-polyisoprene copolymers. Polym Chem 2014;5:3379-88.
[295] Jiang F, Zhang Y, Fang C, Wang Z, Wang Z. From soft to strong elastomers: the role of
additional crosslinkings in copolymer-grafted multiwalled carbon nanotube composite
thermoplastic elastomers. RSC Adv 2014;4:60079-85.
[296] Jiang F, Zhang Y, Wang Z, Fang H, Ding Y, Xu H, et al. Synthesis and characterization
of nanostructured copolymer-grafted multiwalled carbon nanotube composite
thermoplastic elastomers toward unique morphology and strongly enhanced mechanical
properties. Ind Eng Chem Res 2014;53:20154-67.
96

[297] Yu J, Wang J, Wang C, Liu Y, Xu Y, Tang C, Chu F. UV-Absorbent Lignin-Based


Multi-Arm Star Thermoplastic Elastomers. Macromol Rapid Commun 2015;36:398-
404.
[298] Tran CD, Chen J, Keum JK, Naskar AK. A new class of renewable thermoplastics with
extraordinary performance from nanostructured lignin-elastomers. Adv Funct Mater
2016;26:2677-85.
[299] Huang Y, Zheng Y, Sarkar A, Xu Y, Stefik M, Benicewicz BC. Matrix-Free Polymer
Nanocomposite Thermoplastic Elastomers. Macromolecules 2017;50:4742-53.
[300] Jiang F, Zhang Y, Wang Z, Wang W, Xu Z, Wang Z. Combination of magnetic and
enhanced mechanical properties for copolymer-grafted magnetite composite
thermoplastic elastomers. ACS Appl Mater Interfaces 2015;7:10563-75.
[301] Zhang J, Li T, Mannion AM, Schneiderman DK, Hillmyer MA, Bates FS. Tough and
sustainable graft block copolymer thermoplastics. ACS Macro Lett 2016;5:407-12.
[302] Sheiko SS, Sumerlin BS, Matyjaszewski K. Cylindrical molecular brushes: Synthesis,
characterization, and properties. Prog Polym Sci 2008;33:759-85.
[303] Cao Z, Carrillo JMY, Sheiko SS, Dobrynin AV. Computer simulations of bottle
brushes: From melts to soft networks. Macromolecules 2015;48:5006-15.
[304] Paturej Jł, Sheiko SS, Panyukov S, Rubinstein M. Molecular structure of bottlebrush
polymers in melts. Sci Adv 2016;2:e1601478/1-13.
[305] Liang H, Cao Z, Wang Z, Sheiko SS, Dobrynin AV. Combs and bottlebrushes in a melt.
Macromolecules 2017;50:3430-7.
[306] Pakula T, Zhang Y, Matyjaszewski K, Lee HI, Boerner H, Qin S, et al. Molecular
brushes as super-soft elastomers. Polymer 2006;47:7198-206.
[307] Daniel WFM, Burdyńska J, Vatankhah-Varnoosfaderani M, Matyjaszewski K, Paturej
J, Rubinstein M, et al. Solvent-free, supersoft and superelastic bottlebrush melts and
networks. Nat Mater 2016;15:183-90.
[308] Vatankhah-Varnosfaderani M, Daniel WFM, Everhart MH, Pandya AA, Liang H,
Matyjaszewski K, et al. Mimicking biological stress–strain behaviour with synthetic
elastomers. Nature 2017;549:497-501.
97

[309] Vatankhah-Varnosfaderani M, Keith AN, Cong Y, Liang H, Rosenthal M, Sztucki M, et


al. Chameleon-like elastomers with molecularly encoded strain-adaptive stiffening and
coloration. Science 2018;359:1509-13.
[310] Voit BI, Lederer A. Hyperbranched and highly branched polymer architectures s
synthetic strategies and major characterization aspects. Chem Rev 2009;109:5924-73.
[311] Hutchings LR, Dodds JM, Roberts-Bleming SJ. Hypermacs: Highly branched polymers
prepared by the polycondensation of AB2 macromonomers, synthesis and
characterization. Macromolecules 2005;38:5970-80.
[312] Hutchings LR, Dodds JM, Rees D, Kimani SM, Wu JJ, Smith E. Hypermacs to
hyperblocks: A novel class of branched thermoplastic elastomer. Macromolecules
2009;42:8675-87.
[313] Hutchings LR, Agostini S, Hamley IW, Hermida-Merino D. Chain architecture as an
orthogonal parameter to influence block copolymer morphology. Synthesis and
characterization of hyperbranched block copolymers: Hyperblocks. Macromolecules
2015;48:8806-22.
[314] Nugay N, Nugay T, Deodhar T, Keszler BL, Kennedy JP. Low cost bifunctional
initiators for bidirectional living cationic polymerization of olefins. II. hyperbranched
styrene–isobutylene–styrene triblocks with superior combination of properties. J Polym
Sci Part A Polym Chem 2018;56:705-13.
[315] 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
polybutadienes. Macromolecules 2002;35:6592-7.
[316] Knauss DM, Huang T. Star-block-linear-block-star triblock (pom-pom) polystyrene by
convergent living anionic polymerization. Macromolecules 2002;35:2055-62.
[317] Huang T, Knauss DM. Synthesis and characterization of (star po1ystyrene)-block-(linear
polydimethylsi1oxane)-block-(star polystyrene) triblock copolymers. Polym Bull
2002;49:143-50.
[318] Huang T, Knauss DM. (Star PS)-block-(Linear PI)-block-(Star PS) triblock copolymers
- thermoplastic elastomers with complex branchd architectures. Macromol Symp
2004;215:81-93.
98

[319] Wijayasekara DB, Huang T, Richardson JM, Knauss DM, Bailey TS. The role of

architecture in the melt-state self-assembly of (polystyrene) star - b‑ (polyisoprene)

linear - b‑ (polystyrene) star pom-pom triblock copolymers. Macromolecules

2016;49:595-608.
[320] Higashihara T, Faust R, Inoue K, Hirao A. Synthesis of novel multifunctional
polyisobutylenes at chain end(s) and their application to AnB asymmetric star and
AnBAn pompom polymers by combination of living cationic and anionic
polymerizations. Macromolecules 2008;41:5616-25.
[321] Miura Y, Satoh K, Kamigaito M, Okamoto Y, Kaneko T, Jinnai H, et al. AxBAx-type
block-graft polymers with middle soft segments and outer hard graft chains by
ruthenium-catalyzed living radical polymerization: Synthesis and characterization.
Macromolecules 2007;40:465-73.
[322] Miura Y, Kaneko T, Satoh K, Kamigaito M, Jinnai H, Okamoto Y. AxBAx‐type
block–graft polymers with soft methacrylate middle segments and hard styrene outer
grafts: Synthesis, morphology, and mechanical properties. Chem Asian J 2007;2:662-72.
[323] Greef TFAd, Meijer EW. Supramolecular polymers. Nature 2008;453:171-3.
[324] Yoon J, Lee HJ, Stafford CM. Thermoplastic elastomers based on ionic liquid and
poly(vinyl alcohol). Macromolecules 2011;44: 2170-8.
[325] Wang Y, Liu X, Li S, Li T, Song Y, Li Z, et al. Transparent, healable elastomers with
high mechanical strength and elasticity derived from hydrogen-bonded polymer
complexes. ACS Appl Mater Interfaces 2017;9:29120-9.
[326] Scavuzzo J, Tomita S, Cheng S, Liu H, Gao M, Kennedy JP, et al. Supramolecular
elastomers: Self-assembling star-blocks of soft polyisobutylene and hard oligo( β-
alanine) segments. Macromolecules 2015;48:1077-86.
[327] Luo MC, Zeng J, Xie ZT, Wei LY, Huang G, Wu J. Impact of hydrogen bonds
dynamics on mechanical behavior of supramolecular elastomer. Polymer 2016;105:221-
6.
[328] Cordier P, Tournilhac F, Soulié-Ziakovic C, Leibler L. Self-healing and
thermoreversible rubber from supramolecular assembly. Nature 2008;451:977-80.
99

[329] Mather BD, Baker MB, Beyer FL, Berg MAG, Green MD, Long TE. Supramolecular
triblock copolymers containing complementary nucleobase molecular recognition.
Macromolecules 2007;40:6834-45.
[330] Mather BD, Baker MB, Beyer FL, Green MD, Berg MAG, Long TE. Multiple
hydrogen bonding for the noncovalent attachment of ionic functionality in triblock
copolymers. Macromolecules 2007;40:4396-8.
[331] Hayashi M, Matsushima S, Noro A, Matsushita Y. Mechanical property enhancement
of aba block copolymer-based elastomers by incorporating transient cross-links into soft
middle block. Macromolecules 2015;48:421-31.
[332] Hayashi M, Noro A, Matsushita Y. Highly extensible supramolecular elastomers with
large stress generation capability originating from multiple hydrogen bonds on the long
soft network strands. Macromol Rapid Comm 2016;37:678-84.
[333] Chen Y, Kushner AM, Williams GA, Guan Z. Multiphase design of autonomic self-
healing thermoplastic elastomers. Nat Chem 2012;4:467-72.
[334] Cheng S, Beyer FL, Mather BD, Moore RB, Long TE. Phosphonium-containing ABA
triblock copolymers: Controlled free radical polymerization of phosphonium ionic
liquids. Macromolecules 2011;44:6509-17.
[335] Voorhaar L, Diaz MM, Leroux F, Rogers S, Abakumov AM, Tendeloo GV, et al.
Supramolecular thermoplastics and thermoplastic elastomer materials with self-healing
ability based on oligomeric charged triblock copolymers. NPG Asia Mater
2017;9:e385/1-10.
[336] Jiang F, Fang C, Zhang J, Wang W, Wang Z. Triblock copolymer elastomers with
enhanced mechanical properties synthesized by raft polymerization and subsequent
quaternization through incorporation of a comonomer with imidazole groups of about
2.0 mass percentage. Macromolecules 2017;50:6218-26.
[337] Wang W, Wang X, Jiang F, Wang Z. Synthesis, order-to-disorder transition,
microphase morphology and mechanical properties of BAB triblock copolymer
elastomers with hard middle block and soft outer blocks. Polym Chem 2018;9:3067-79.
[338] Rungswang W, Kotaki M, Shimojima T, Kimura G, Sakurai S, Chirachanchai S.
Directing thermoplastic elastomer microdomain parallel to fiber axis: A model case of
sebs with benzoxazine through π-π stacking. Macromolecules 2011;44:9276-85.
100

[339] Xing W, Li H, Huang G, Cai LH, Wu J. Graphene oxide induced crosslinking and
reinforcement of elastomers. Compos Sci Technol 2017;144:223-9.
[340] Araby S, Zaman I, Meng Q, Kawashima N, Michelmore A, Kuan HC, Majewski P, Ma
J. Melt compounding with graphene to develop functional, high-performance
elastomers. Nanotechnology 2013;24:165601/1-5.
[341] Araby S, Zhang L, Kuan HC, Dai JB, Majewski P, Ma J. A novel approach to
electrically and thermally conductive elastomers using graphene. Polymer
2013;54:3663-70.
[342] Wang C, Liu N, Allen R, Tok J, Wu Y, Zhang F, et al. A rapid and efficient self‐
healing thermo‐reversible elastomer crosslinked with graphene oxide. Adv Mater
2013;25:5785-90.
[343] Burattini S, Greenland B, Merino D, Weng W, Seppala J, Colquhoun H, et al. A
healable supramolecular polymer blend based on aromatic π−π stacking and hydrogen-
bonding interactions. J Am Chem Soc 2010;132:12051-8.
[344] Burattini S, Greenland B, Hayes W, Mackay M, Rowan S, Colquhoun H. A
supramolecular polymer based on tweezer-type π−π stacking interactions: Molecular
design for healability and enhanced toughness. Chem Mater 2011;23:6-8.
[345] Mozhdehi D, Ayala S, Cromwell OR, Guan Z. Self-healing multiphase polymers via
dynamic metal–ligand interactions. J Am Chem Soc 2014;136:16128-31.
[346] Liu X, Zhao RY, Zhao TP, Liu CY, Yang S, Chen EQ. An ABA triblock containing a
central soft block of poly [2, 5-di(n-hexogycarbonyl) styrene]and outer hard block of
poly (4-vinylpyridine): synthesis, phase behavior and mechanical enhancement. RSC
Adv 2014;4:18431-41.
[347] Ibarra L, Rodríguez A, Mora-Barrantes I. Crosslinking of carboxylated nitrile rubber
(XNBR) induced by coordination with anhydrous copper sulfate. Polym Int
2009;58:218-26.
[348] Xie F, Huang C, Wang F, Huang L, Weiss RA, Leng J, Liu Y. Carboxyl-Terminated
Polybutadiene–Poly (styrene-co-4-vinylpyridine) Supramolecular Thermoplastic
Elastomers and Their Shape Memory Behavior. Macromolecules 2016;49:7322-30.
[349] Burnworth M, Tang L, Kumpfer JR, Duncan AJ, Beyer FL, Fiore GL, et al. Optically
healable supramolecular polymers. Nature 2011;472:334-7.
101

[350] Kumpfer JR, Jin J, Rowan SJ. Stimuli-responsive europium-containing metallo-


supramolecular polymers. J Mater Chem 2010;20:145-51.
[351] Chen S, Mahmood N, Beiner M, Binder WH. Self‐healing materials from V‐and
H‐shaped supramolecular architectures. Angew Chem Int Ed 2015;54:10188-92.
[352] Anonymous. Thermoplastic Elastomers Market by Type, End-use Industry, Region -
Global Forecast to 2022. https://www.marketsandmarkets.com/Market-
Reports/thermoplastic-elastomers-market-1012.html 2017;Accessed Sep 2018
[353] Jiang J, Ye G, Wang Z, Lu Y, Chen J, Matyjaszewski K. Heteroatom-Doped Carbon
Dots (CDs) as a New Class of Metal-Free Photocatalysts for PET-RAFT Polymerization
under Visible Light and Sunlight. Angew Chem Int Ed 2018;57:12037-42.
[354] Dadashi-Silab S, Doran S, Yagci Y. Photoinduced Electron Transfer Reactions for
Macromolecular Syntheses. Chem Rev 2016;116:10212-75.
[355] Lorandi F, Wang Y, Fantin M, Matyjaszewski K. Ab Initio Emulsion Atom-Transfer
Radical Polymerization. Angew Chem Int Ed 2018;57:8270-4.
[356] Jakubowski W. Adapting Atom Transfer Radical Polymerization to Industrial Scale
Production: The Ultimate ATRPSMTechnology. Prog Controlled Radical
Polymerization: Mechan Techniq: American Chemical Society; 2012. p. 203-16.
[357] Brzytwa AJ, Johnston J. Scaled production of RAFT CTA - a star performer. Polym Prepr
Am Chem Soc Div Polym Chem 2011;52(2):533-4.
[358] Destarac M. Industrial development of reversible-deactivation radical polymerization: is
the induction period over? Polym Chem 2018;9:4947-67.
[359] Zhu Y, Romain C, Williams CK. Sustainable polymers from renewable resources.
Nature 2016;540:354-62.
[360] Schneiderman DK, Hillmyer MA. 50th anniversary perspective: There is a great future
in sustainable polymers. Macromolecules 2017;50:3733-49

You might also like