P12_S_Bates_Multiblock Polymers - Panacea or Pandora's Box

Embed Size (px)

Citation preview

  • 8/11/2019 P12_S_Bates_Multiblock Polymers - Panacea or Pandora's Box

    1/8

    DOI: 10.1126/science.1215368, 434 (2012);336Science

    et al.Frank S. BatesMultiblock Polymers: Panacea or Pandora's Box?

    This copy is for your personal, non-commercial use only.

    clicking here.colleagues, clients, or customers by, you can order high-quality copies for yourIf you wish to distribute this article to others

    here.following the guidelines

    can be obtained byPermission to republish or repurpose articles or portions of articles

    ):June 15, 2012www.sciencemag.org (this information is current as of

    The following resources related to this article are available online at

    http://www.sciencemag.org/content/336/6080/434.full.htmlversion of this article at:

    including high-resolution figures, can be found in the onlineUpdated information and services,

    http://www.sciencemag.org/content/suppl/2012/04/25/336.6080.434.DC1.htmlcan be found at:Supporting Online Material

    http://www.sciencemag.org/content/336/6080/434.full.html#ref-list-1, 6 of which can be accessed free:cites 48 articlesThis article

    http://www.sciencemag.org/cgi/collection/mat_sci

    Materials Sciencesubject collections:This article appears in the following

    registered trademark of AAAS.is aScience2012 by the American Association for the Advancement of Science; all rights reserved. The title

    CopyrighAmerican Association for the Advancement of Science, 1200 New York Avenue NW, Washington, DC 20005.(print ISSN 0036-8075; online ISSN 1095-9203) is published weekly, except the last week in December, by thScience

    http://www.sciencemag.org/about/permissions.dtlhttp://www.sciencemag.org/about/permissions.dtlhttp://www.sciencemag.org/about/permissions.dtlhttp://www.sciencemag.org/about/permissions.dtlhttp://www.sciencemag.org/about/permissions.dtlhttp://www.sciencemag.org/about/permissions.dtlhttp://www.sciencemag.org/content/336/6080/434.full.htmlhttp://www.sciencemag.org/content/336/6080/434.full.htmlhttp://www.sciencemag.org/content/336/6080/434.full.html#ref-list-1http://www.sciencemag.org/content/336/6080/434.full.html#ref-list-1http://www.sciencemag.org/content/336/6080/434.full.html#ref-list-1http://www.sciencemag.org/content/336/6080/434.full.html#ref-list-1http://www.sciencemag.org/cgi/collection/mat_scihttp://www.sciencemag.org/cgi/collection/mat_scihttp://www.sciencemag.org/cgi/collection/mat_scihttp://www.sciencemag.org/content/336/6080/434.full.html#ref-list-1http://www.sciencemag.org/content/336/6080/434.full.htmlhttp://www.sciencemag.org/about/permissions.dtlhttp://www.sciencemag.org/about/permissions.dtlhttp://oascentral.sciencemag.org/RealMedia/ads/click_lx.ads/sciencemag/cgi/reprint/L22/60419773/Top1/AAAS/PDF-Sigma-Science-120101/sh4971_Science_sponsor_logo_v5.raw/71304a356c552f6343425941424c4a44?x
  • 8/11/2019 P12_S_Bates_Multiblock Polymers - Panacea or Pandora's Box

    2/8

    Multiblock Polymers:Panacea or Pandoras Box?Frank S. Bates,1* Marc A. Hillmyer,2 Timothy P. Lodge,1,2 Christopher M. Bates,3

    Kris T. Delaney,

    4

    Glenn H. Fredrickson

    4,5

    Advances in synthetic polymer chemistry have unleashed seemingly unlimited strategies for producingblock polymers with arbitrary numbers (n) and types (k) of unique sequences of repeating units.Increasing (k,n) leads to a geometric expansion of possible molecular architectures, beyondconventional ABA-type triblock copolymers (k= 2,n = 3), offering alluring opportunities to generateexquisitely tailored materials with unparalleled control over nanoscale-domain geometry, packingsymmetry, and chemical composition. Transforming this potential into targeted structures endowedwith useful properties hinges on imaginative molecular designs guided by predictive theory andcomputer simulation. Here, we review recent developments in the field of block polymers.

    Block polymers, hybrid macromolecules

    constructed by linking together discrete

    linear chains comprising dozens to hun-

    dreds of chemically identical repeating units,spontaneously assemble into exquisitely ordered

    soft materials (1). Precise synthesis of these self-

    assembling compounds offers extraordinary con-

    trol over the resulting morphology, spanning length

    scales from less than a few nanometers to sev-

    eral micrometers, enabling a diverse and expand-

    ing range of practical applications in, for example,

    drug delivery (2), microelectronic materials (3),

    and advanced plastics (4). Yet, only a small sub-

    set of the vast array of feasible molecular ar-

    chitectures has been explored, and just a handful

    of these compounds have been developed into

    commercial products. Combining more chemi-

    cally distinct blocks and block types, beyond theestablished AB and ABA diblock and triblock

    copolymers, offers unparalleled opportunities for

    designing new nanostructured materials with en-

    hanced functionality and properties, often with-

    out adding substantially to the cost of production.

    But how do we choose among the myriad mo-

    lecular design possibilities?

    Modern synthetic methods afford access to

    a broad portfolio of multiblock molecular archi-

    tectures, as illustrated schematically in Fig. 1.

    (Although there is no universal definition of the

    minimum molar mass that constitutes a polymer

    block, generally 10 to 20 monomer repeat units,

    and frequently more than 100, are used). Linear

    AB diblock copolymers have been investigated

    most extensively, leading to a comprehensive ex-

    perimental and theoretical understanding of their

    bulk- and solution-phase behavior (1). Exten-

    sion to linear alternating multiblock copolymers

    (ABA, ABABA, etc.) can lead to profound con-sequences on the physical properties (for exam-

    ple, enhanced elasticity and fracture toughness)

    (5) without drastically influencing the associated

    phase behavior. Introduction of a third block

    type, C, dramatically expands the spectrum of

    accessible nanostructured morphologies ormicro-

    phases.Whereas AB and ABA copolymers typ-

    ically adopt four familiar microphase structures

    (lamellae, double gyroid, cylinders, and spheres),

    many more ordered phases have been documented

    with ABC triblock terpolymers (1,69). Adding

    additional numbers of blocks (n) and chemi-

    cally distinct block types (k) rapidly expands the

    number of unique sequences (see Box 1), eachcapable of producing a host of nanostructures.

    Cyclic and branched architectures further ex-

    pand the possibilities (10), leading to dizzyi

    phase complexity.

    Block polymer phase behavior is determin

    by a suite of molecular variables, in additi

    to the molecular topology and specific block

    quences, as summarized in Table 1. Primary f

    tors include the degree of polymerization of ea

    block, Ni, and the associated binary segme

    segment interaction parameters, cij (where i a

    j refer to chemically distinct repeat units) (Box 2). (Alternatively, the composition fi = N

    and overall molecular size N= SNican be usin place of defining each block length, wher

    common segment volume defines Ni). Secon

    ary factors (not discussed further) include blo

    flexibility (for instance, stiff versus flexible chain

    the distribution in block lengths (dispersity), a

    any additional sub-block structure such as

    ternating, random, or tapered sequences of

    peat units. Obviously, increasing n and/o

    even slightly, compounded by scores of prac

    cal options in choosing the block chemistr

    results in an expansive parameter space an

    REVIEW

    1Department of Chemical Engineering and Materials Science,University of Minnesota, Minneapolis, MN 55455, USA. 2De-partment of Chemistry, University of Minnesota, Minneapolis,MN 55455, USA. 3Department of Chemistry, The University ofTexas at Austin, Austin, TX 78712, USA. 4Materials ResearchLaboratory, University of California, Santa Barbara, CA 93106,USA. 5Department of Chemical Engineering, University of Cal-ifornia, Santa Barbara, CA 93106, USA.

    *To whom correspondence should be addressed. E-mail:[email protected]

    Table 1. Molecular variables that influenblock polymer self-assembly.

    Primary

    Topology; linear, branched, etc.

    Number of blocks, n

    Number of block types, k

    Block degree of polymerization, N iInteraction parameters, c ij

    Secondary

    Block flexibility, b iMolar mass distribution, i

    Sub-block structure (e.g., tapered)Composition heterogeneity

    1

    11

    2

    4

    23

    3

    1

    Fig. 1.A subset of the vast structural complexity with two (k= 2) or three (k= 3) block types producby varying the number of blocks (n) and the functionality of the connector at each block-block junctu(difunctional, circles; trifunctional, triangles). The number and type of connectors needed are given each structure.

    27 APRIL 2012 VOL 336 SCIENCE www.sciencemag.org34

  • 8/11/2019 P12_S_Bates_Multiblock Polymers - Panacea or Pandora's Box

    3/8

    boundless array of possible structures and che

    ical functionalities.

    Multiblock polymers, such as those sketch

    in Fig. 1, offer unlimited potential for designi

    soft materials with precisely specified structur

    subject to two critical limitations: (i) The stru

    tures can be accessed synthetically, and (ii) p

    dictive theoretical tools to guide molecular des

    are available. A parallel can be drawn with t

    complexity and diverse functions associated w

    proteins, another well-known class of multifutional macromolecules. Theorists strive to pred

    polypeptide structure and ultimate function ba

    on models that employ quantum chemical a

    statistical mechanical tools capable of account

    for both short-range (~0.1 nm) and long-ran

    interactions mediated by an aqueous environm

    and constrained by the macromolecular arc

    tecture (11). Essentially any sequence of less th

    100 residues of the 20 naturally occurring ami

    acids (and a bevy of unnatural variants) can

    produced using commercially available so

    phase peptide synthesizers. The challenge is

    anticipate which linear combinations of am

    acids, drawn from 20n

    possibilities, will leadthe desired secondary, tertiary, and even quat

    nary structures and corresponding functions.

    In some respects, block polymers appear

    pose a less daunting theoretical challenge, p

    ticularly for flexible and noncrystalline polyme

    because the chain structure at the monomer le

    does not play a primary role in determining

    domain geometry or packing symmetry. Mo

    over, with no direct analog to solid-phase pept

    synthesis, contemporary research focuses on p

    ducing block polymers with a minimum deg

    of complexity (that is, the smallestn andk) n

    essary to achieve a desired morphology and co

    plement of properties. However, whereas protfolding is intrinsically intramolecular, supram

    lecular self-assembly of block polymers lea

    to ordered structures with unit cells that may co

    tain many thousands of polymer chains (m

    lions of atoms), resulting in unique theoreti

    and computational challenges. Hence, even

    seemingly primitive case of ABC triblocks (k=

    n = 3) is only marginally understood in ter

    of the resulting phase structure and associat

    properties.

    Synthesis

    Block polymer synthesis has evolved consid

    ably since the introduction and development

    living anionic polymerization by Szwarc m

    than 50 years ago (12). Numerous strategies t

    enable the covalent connection of two or m

    polymer blocks have been developed, includ

    sequential addition of distinct monomers to

    active polymer chain, chain-coupling strategi

    and transformations of polymer end groups

    accommodate diverse and often incompati

    polymerization mechanisms. Chemical conn

    tors that link together more than two chains

    increasingly available as evidenced by the her

    syntheses of a 31-arm starblock pentapolym

    Box 1. Mnages en blocs.The challenge of enumerating possible block polymer structures is already evident in Fig. 1,

    where it is seen that a substantial number of molecules can be constructed by linking multipleblock species using difunctional or trifunctional linkers. An appreciation for the numbers involvedcan be obtained by considering just the set of linear molecules composed of n blocks, selectedfrom kdifferent species, and connected into a chain by means of n 1 difunctional linkers. Theenumeration of all such linear molecules, subject to the restrictions (i) that adjacent blocks alongeach chain are of different species, (ii) only topologically distinct sequences are counted (i.e., ABis not distinguishable from BA), and (iii) allkblock species appear, is a problem in combinatorialanalysismnages en blocsthat is taken up in the supplementary material.

    Mnages en blocs has close connections to classic problems in permutations with restrictions,including the 19th century problme des mnages(47,48), which asks for the number of waysof seatingn married couples at a circular table, men and women in alternate positions, no wifenext to her husband. The blocs problem is also related to problems involving enumeration ofprotein sequences (49, 50), although in the latter there are normally no species restrictions onadjacent amino acid residues. The restriction (i) for multiblock polymers amounts to the as-sumption that blocks do not acquire size dispersity by virtue of sequence. Though this assumptionis violated for some classes of block polymers formed by statistical linking of monomers and/orpreformed blocks by various (e.g., condensation) chemistries, such nonideal systems are beyondthe scope of the present discussion.

    The object of interest in mnages en blocs isZ(k,n), the number of linear block polymers thatcan be formed from n blocks andkdifferent block species, and is subject to the restrictions (i) to(iii) stated above. This object can be constructed by exclusion from a simpler function Q(k,n) thatis defined as the number of linear block polymers that can be formed fromnblocks selected from

    kspecies and subject only to the restrictions (i) and (ii). An expression for Q is derived in thesupplementary material

    Q(k, n) k2(k 1)

    (n1) n evenk2(k 1)

    (n1)/2[1+ (k 1)(n1)/2] n odd

    Imposing the final restriction (iii) that all kblock species must appear in each molecule is trivialfor the case ofk= 2, because restriction (i) implies restriction (iii); hence, Z(2,n) =Q(2,n). Fork> 2,the principle of exclusion is used to recursively generate Z(k,n) fromQ(k,n)

    Z(k, n) = Q(k,n) k1

    j=2

    kj

    Z(j, n)

    In words, block polymers containingk 1 or fewer block species are removed from the setcomprising Q(k,n) to yield Z(k,n), the subset of molecules in Q(k,n) in which all k blockspecies appear.

    Application of the above formulas leads to notable growth in block polymer diversity withincreasing kand n. The table below summarizes the Z(k,n) function for block enumerationswith k,n 10. One observes that there are no molecules with k > n, as is required byrestriction (iii). For k= 2, there is an odd-even effect with increasing n (e.g., the mnage trois ABA and BAB at n = 3, ABAB at n = 4). Along the diagonal, block polymer complexityexplodes in a factorial manner with Z(n,n) = n!/2.

    Mnages en blocs function Z(k,n)

    Curiously, the maximum ofZ(k,n) for n > 4 is above the diagonal, located atZ(n 1,n)for5 n 8,and atZ(n 2,n) forn= 9, 10. Synthesis of all these distinct linear block polymer molecules willsurely keep synthetic polymer chemists busy for some time!

    k/n 2 3 4 5 6 7 8 9 10

    2 1 2 1 2 1 2 1 2 1

    3 0 3 9 24 45 102 189 402 765

    4 0 0 12 72 300 1092 3612 11,664 36,3005 0 0 0 60 600 3900 21,000 102,120 466,200

    6 0 0 0 0 360 5400 50,400 378,000 2,502,360

    7 0 0 0 0 0 2520 52,920 670,320 6,667,920

    8 0 0 0 0 0 0 20,160 564,480 9,313,920

    9 0 0 0 0 0 0 0 181,440 6,531,840

    10 0 0 0 0 0 0 0 0 1,814,400

    www.sciencemag.org SCIENCE VOL 336 27 APRIL 2012

    RE

  • 8/11/2019 P12_S_Bates_Multiblock Polymers - Panacea or Pandora's Box

    4/8

    (k = 5, n = 31) (13). In principle, all of t

    structures depicted in Fig. 1 could be prepa

    with arbitrary choice of block constituents usi

    modern polymerization methods and polym

    functionalization strategies, including numero

    controlled radical, ring-opening, metal-catalyz

    and ionic polymerizations, often in combinat

    with highly selective and efficient functional-gro

    transformation techniques (14, 15).

    Modern block polymer molecular design

    driven by the intended applications. For exaple, a thermally stable and mechanically rob

    membrane endowed with a dense array of sp

    cifically sized nanopores necessitates a combin

    tion of glassy, ductile, and chemically etchab

    (or cleavable) blocks (16). Biocompatibility f

    ther limits the possible ingredients. The asso

    ated segment-segment interaction parameters a

    optimal block architecture demand a carefu

    targeted synthetic strategy, possibly includ

    multiple routes to the desired structure, draw

    from an ever-expanding synthetic tool kit. A

    though high degrees of molecular uniformity

    generally desirable, in most circumstances so

    level of imperfection (e.g., less than n blocand modest chain-length distribution can be t

    erated [in fact, tailored dispersity represent

    strategic design parameter (17, 18)]; the goa

    to build the appropriate structure in the m

    synthetically economical way possible.

    Figure 2 illustrates routes to all possible

    quences of a linear ABC terpolymer (k= 3,n=

    designed to incorporate three complement

    physical and chemical properties: (i) glassy p

    (styrene) (PS), (ii) rubbery poly(isoprene) (P

    and (iii) chemically etchable and biorenewa

    poly(lactide) (PLA). Although only two of th

    three structures have been reported, PI-PLA-

    is certainly tractable on the basis of a relatprecedent (19).

    The synthetic strategies depicted in Fig. 2

    quire various combinations of ring-opening,

    ionic, and controlled radical polymerizations w

    concomitant end-group manipulations. In fa

    the drive to largern and k will almost certain

    necessitate multiple polymerization mechanis

    and functional-group manipulations, the or

    of which will depend on the desired block

    quence. Complexity in ABC terpolymer syste

    is further increased by introducing the pos

    bility of cyclic motifs and three-point junctio

    (miktoarm stars or brushes). Putting all the pie

    together requires the continued development

    efficient processes and realizable retrosynthe

    analyses.

    Introduction of a fourth block with only th

    monomers (k= 3,n = 4) provides an added le

    of complexity not realized in ABC terpolyme

    Asymmetric ABCA tetrablock terpolym

    (NA NA) offer unique opportunities for d

    signing bulk materials (highlighted in the Str

    ture section) and tailored solution structures (2

    However, independent control over the length

    all four blocks requires synthetic strategies t

    depend on the specific block chemistries. F

    10-4 10-3 10-2 10-1 100 101

    D D

    D

    D

    D

    DDD

    OO

    O

    O

    O

    N

    SO3-

    Estimated with respect to polystyrene

    Box 2. Complexity with c.Discussions of polymer phase behaviorwhether block polymers, blends, or solutionsare

    usually couched in terms of the binary interaction parameter cAB, or simply c. Conceptually, c isa dimensionless measure of the energetic cost of exchanging a repeat unit of polymer A for anequal volume (Vref) unit of polymer B: c = zDw/kT, where kT is the thermal energy (k is theBoltzmann constant, and T is temperature), Dw is the difference between the A/B interactionenergywABand the average of the self-interaction energieswAAandwBB, andzis the number ofnearest neighbors. On the assumption that there is entirely random mixing of A and B units, withentropy arising only from the ideal combinatorial of mixingDSm

    id = kS(fi/Ni)lnfi(wherefiand

    Niare the volume fraction and degree of polymerization of species i, respectively), the Flory-Huggins (or Bragg-Williams or mean-field or regular-solution) free energy of mixing for A and Bpolymers becomes

    DGm = TDSmid + ckTfAfB

    However, in experimental practice, c is imbued with whatever attributes are needed todescribe the data; as such, it actually represents an excess free energy of mixing DGm

    ex thatexhibits both enthalpic and entropic contributions. Therefore, in practice c is recast as c = cH+cS= a/T+ b (wherea andb are empirical parameters), and either a or b may also depend onfand N. This complexity can lead to much confusion when, for example, comparing differentmeasurements on the same system. Furthermore, either a orb may be positive or negative insign. The case where a > 0 andb < 0 leads to the classicalupper critical solution temperaturephase diagram, whereas a < 0 and b > 0 leads to a lower critical solution temperature (i.e.,

    phase separation on heating); both situations are common. Because DSmid

    varies inversely withN,it is negligible for typical molecular weights; consequently, the excess free energy is not just acorrection, but the main contribution.

    The characteristic magnitude of cABspans several orders of magnitude, as illustrated below.From a predictive point of view, the situation is quite unsatisfactory. For systems in which shortrange, isotropic, dispersive interactions dominate, the solubility parameter approach might beexpected to yield reasonable estimates [i.e., c cH = Vref/kT(dA dB)

    2, where dA and dB arethe associated segment solubility parameters]; however, in practice, agreement with experi-ment is often remarkably poor. Additionally, although solubility parameters can be used toanticipate qualitatively where a particular blend might lie on this continuum, there is no generaltheoretical approach for predicting the relative contributions ofcHandcSor for computing eitherterm to within, say, a factor of 2. As phase boundaries in polymer systems typically depend on thedimensionless product cN, it is therefore not possible to even anticipate the molecular weightsneeded to produce an experimentally accessible phase boundary.

    Blends of hydrogenous and deuterated isotopomers represent the simplest case. Here, it is

    usually found thatc 103

    , andcHcan be understood in terms of differences in zero point energyand the polarizabilities of CH and CD bonds. Nevertheless, cSis not negligible, for reasons thatremain elusive. For example, for (H/D) poly(styrene) blends, c(373 K) 0.00014, but cSis ofcomparable magnitude to cHand is negative (51). For poly(styrene)/poly(methylmethacrylate), apair of prevalent thermoplastics, c (373 K) 0.038, a relatively modest value; however, cS ispositive and, surprisingly, represents ~70% of the total, so that c itself has almost no Tdepen-dence (52). When polar or ionic polymers are mixed with nonpolar partners, c can exceed unity.Here, the fundamental interactions are relatively strong, long-ranged, anisotropic, and/or non-pairwise additive, yet entropic contributions still play a substantial role. For example, forpoly(styrene)/poly(lactide), c(373 K) 0.15, but the magnitude of cSat this temperature isabout half that of cH (53).

    27 APRIL 2012 VOL 336 SCIENCE www.sciencemag.org36

    REVIEW

  • 8/11/2019 P12_S_Bates_Multiblock Polymers - Panacea or Pandora's Box

    5/8

    example, the addition of water-soluble and bio-compatible poly(ethylene oxide) (PEO) to both

    ends of a PS-PI core should be feasible. Se-

    quential anionic polymerization of styrene and

    isoprene initiated with an alkyllithium reagent

    containing a protected hydroxyl group (21) fol-

    lowed by end-capping with one unit of ethylene

    oxide generates a heterotelechelic diblock co-

    polymer. Activation of the free alcohol end en-

    ables the ring-opening polymerization of ethylene

    oxide, without compromising the protecting group

    at the other chain end, leading to a linear BCA

    triblock; the terminal end of the A block must

    be capped to prohibit further growth during sub-

    sequent polymerization of ethylene oxide fromthe B segment in this scheme. Unmasking and

    activation of the hydroxyl group at the B ter-

    minus allows for independent control of NArelative to NA. Such asymmetric ABCA tetra-

    block polymers have not been reported, but this

    approach could enable remarkable tunability of

    ordered microstructures (see below).

    New organocatalytic polymerization methods

    (22), controlled synthesis of conducting poly-

    mers (23), and various clickstrategies (14), all

    developed over the past decade, exemplify how

    advances in chemistry have allowed access to

    hybrid structures that simply could not be made

    before. However, deciding what macromolecular

    masterpieceto synthesize using the expanding

    palette of monomer paints and polymer syn-

    thesis brushesrepresents a growing conundrum.

    Targeting new block structures for purely aesthetic

    reasons is impractical. Today, synthetic polymer

    chemists can prepare nearly any architecture

    with any set of desired chemistries, for com-

    modity and value-added applications, paralleling

    the synthesis of small molecules for therapeutic

    markets. Organic chemists can build complicated

    small-molecule structures with an amazing, al-

    most arbitrary array of chemical functionalities;

    to know which molecule is efficacious requiresa better and more complete understanding of

    biological action. Analogously, given the heavy

    investment required for creating even a single

    new multiblock structure (24), advances in the

    block polymer/soft materials arena require (i) a

    deeper understanding of how block architecture

    influences structure and, thus, properties and (ii)

    advances in predictive theories that can guide

    synthetic chemists.

    Structure

    The fundamental principles governing block

    copolymer self-assembly were established in

    the 1970s, culminating in Helfands strong seg-regation (25) and Leiblers weak segregation

    (26) analyses of AB diblock copolymers. These

    theories have since been subsumed into a com-

    prehensive mean-field framework known as self-

    consistent field theory (SCFT) (see the Theory

    section). Equilibrium-phase behavior represents

    a compromise between minimizing unfavorable

    segment-segment contacts, mediated bycAB, andmaximizing configurational entropy, which is

    inversely proportional toNand quadratically de-

    pendent on the extent of chain stretching relative

    to the unperturbed state. Simplifying assump-

    tions, including Gaussian chain (random walk)

    statistics and constant density, enables tractable

    computational schemes that have accounted for

    all of the experimentally observed diblock mor-

    phologies (27). Increasing n and k introduces

    additional complexity, greatly compounded by

    the choice of block sequences along with the

    other molecular parameters listed in Table 1.

    In the limit of strong segregation, multiblock

    polymers segregate into relatively pure domains

    separated from neighboring domains by as many

    ask(k 1)/2 distinct and narrow interfaces char-

    acterized by interfacial tensiongij~ (cij)1/2; strong

    segregation impliesNi Nj>> 10/cij. Reducing

    the individual values ofcij leads to broader terfaces and, ultimately, mixing ofi andjbloc

    Thus, an ABC triblock terpolymer (k= 3,n =

    may contain one, two, or three types of interfa

    (or none if entirely disordered). For example,

    condition cAB = cBC > 10/N, t

    x= 1 tetrablock terpolymer will exhibit a tw

    domain lamellar morphology with equival

    interfaces separating A and A domains fro

    mixed B/C domains (Fig. 3A). Increasing t

    segregation strength to cAB= cAC= cBC/2 wresult in a three-domain lamellar configurat

    Fig. 2. Synthetic routes to produce the three distinct linear triblockterpolymer structures composed of PS, PI, and PLA blocks, using com-binations of living anionic, metal-catalyzed ring-opening, and revers-

    ible addition-fragmentation chain transfercontrolled radical polymerizatioR and R, alkyl groups; Ph, phenyl; p, q, and r, degrees of polymerizatioEt, ethyl.

    www.sciencemag.org SCIENCE VOL 336 27 APRIL 2012

    RE

  • 8/11/2019 P12_S_Bates_Multiblock Polymers - Panacea or Pandora's Box

    6/8

    with energetically equivalent A/B and C/A in-

    terfaces that exactly balance the internal B/C

    interface (30). Further elevating cBCwill induce

    changes in the domain geometry that minimize

    the overall interfacial free energy (P

    ij

    Aijgij,

    where Aij is the i-j interfacial area) subject to

    entropic penalties associated with packing the

    blocks into the resulting structures. The sequence

    of domains sketched in Fig. 3A, radial and axial

    segregated cylinders followed by a Janus sphere,illustrates phase transitions that capture this ef-

    fect by reducing ABC/(AAC+ AAB), the ratio of

    interfacial surface areas. (We note that this hy-

    pothetical sequence of morphologies may be su-

    perseded by other morphologies and never realized

    in practice). Clearly, addingan Ablock to the ABC

    sequence offers specific control over the shape

    and subdivision of the self-assembled domains.

    Interdomain packing can be considerably in-

    fluenced by the symmetry parameterx, as shownin Fig. 3B. We have selected the axially segre-

    gated cylindrical domain structure to illustrate

    this point. Tetrablock terpolymers (k= 3,n = 4)

    can produce ordered cylinders containing B andC subdomains, even with NA+ NA = N/2 (31);

    cylindrical structures at such low matrix com-

    positions are inaccessible with diblock copoly-

    mers. For x = 1, there will be no interdomain

    axial order because, under the stated conditions,

    the intradomain structure does not influence

    the corona chains that control cylinder packing;

    two-dimensional (2D) hexagonal packing would

    be anticipated. However, for x > 1, the C do-mains will be decorated with A blocks, which

    are longer than the A blocks emanating from

    the B portions of the cylinder. This arrangement

    introduces an effective anisotropic interdomain

    potential that should induce 3D order. The tend-

    ency to pair short and long corona chains rep-

    resents a purely entropic effect, one that minimizes

    unfavorable chain stretching and compression

    at constant density. A different cylinder-packing

    symmetry (for instance, tetragonal) would be re-quired to permit uniform chain packing; neither

    three- nor sixfold symmetry supports complete

    in-plane alternation of C and B domains. Such

    symmetry-breaking also should apply to the Janus

    sphere geometry (and radial segregated cylinders)

    creating dipole-dipole interactions, thus offering

    the fascinating possibility of tailoring domain

    orientation within specific ordered lattices anal-

    ogous to spin alignment in magnetic materials (32)

    [e.g., noncentrosymmetric ferromagnetic order-

    ing (33) of spheres on a body-centered cubic

    (bcc) crystal as shown in Fig. 3B] and pairing of

    colloidal Janus particles (34).

    Simply changing the sequence from ABCA

    to ABAC while holding all other parameters

    constant leads to qualitatively different phase be-

    havior. Fixing cAB = cAC 1, shorchains (emanating from the red B

    mains) will show a preference to pnear long A chains (emanating frthe blue C domains) to maintain costant density while minimizing unfavable chain stretching and compressiThis induces 3D cylindrical order aa ferromagnetic-like arrangementJanus-type spheres.

    BC

    /AB

    A A'

    BC

    AC

    AB

    1

    21

    1

    4

    R

    2d=

    +> > >

    0

    AB=

    AC

    A B C A'B/C

    Noorder

    1>1=

    Order

    2R

    d

    A

    B

    27 APRIL 2012 VOL 336 SCIENCE www.sciencemag.org38

    REVIEW

  • 8/11/2019 P12_S_Bates_Multiblock Polymers - Panacea or Pandora's Box

    7/8

    characterized by effective Hamiltonians (or ac-

    tions) H that are highly nonlinear and nonlocal

    functionals of the field variables and have ex-

    plicit dependence on segmental interaction and

    chain architecture parameters, such as cijandNi.

    The Hamiltonian is generally also a complex-

    valued functional, which can have important con-

    sequences for generating numerical solutions.

    Two major classes of field-based simulations

    can be built on top of the Edwards framework:

    (i) SCFT (39), where mean-field solutions aresought corresponding to saddle points of H,

    and (ii) field-theoretic simulations (FTS), refer-

    ring to stochastic numerical sampling of the full

    (complex-valued) field theory (40). SCFT sim-

    ulations are considerably less expensive than FTS,

    because they seek only a single saddle-point field

    configuration, but SCFT neglects field fluctu-

    ations that are especially important in solvated

    polymer systems. In the case of highmolar mass

    undiluted block polymers, such fluctuation ef-

    fects are substantial only near order-disorder

    phase boundaries and associated critical points,

    so SCFT can be applied to polymer melts with

    relative impunity.Even within the confines of SCFT, however,

    establishing the phase map for a given block

    polymer [for example, from the (k,n) linear mul-

    tiblock family] can be a daunting challenge. First,

    the parameter space is large: Even in the case of

    an ABC triblock, the phase map is defined by

    a complicated surface within the 5D parameter

    spacecABN, cACN, cBCN, fA (= NA/N), and fBthat has yet to be delineated in detail and would

    entail a Herculean effort to complete. Second,

    the mean-field free-energy landscape explored

    in SCFT simulations is rough, so simulations

    launched in large cells from random initial field

    configurations and subject only to periodic bound-

    ary conditions tend to yield defective morphol-

    ogies that are metastable, rather than stable

    (lowest in free energy) for the specified param-

    eters. An example is shown in Fig. 4, where such

    a large-cell SCFT simulation is conducted for

    an ABC triblock melt using parameters (28)that coincide with a PI-PS-PEO system known

    to have a stable orthorhombicFddd(O70) phase

    (7). The simulation launched from random ini-

    tial conditions settles into a metastable, highly

    defective Fdddstructure (Fig. 4A), whereas an

    analogous simulation initialized from a determi-

    nistic seed with the proper symmetry falls quickly

    into the stable O70 phase (Fig. 4B). Occasionally,

    large-cell simulations will also settle into defect-

    free, metastable phases of a different symmetry

    than the stable phase.

    Such challenges of metastability are familiar

    to researchers in a variety of fields that rely on

    global optimization; indeed, practitioners of com-putational protein folding struggle with rough

    energy landscapes (11). Researchers in that field

    have the advantage that nature has apparently

    engineered protein sequences that are resistant

    to misfolding; nonetheless, they lack a quantita-

    tive mean-field theory, and their folding results

    do not enjoy the universality across broad fam-

    ilies of monomers as do SCFT predictions for

    multiblock polymers.

    Large-cell SCFT simulations are the tool

    choice when prospecting for candidate order

    phases in a new polymer system for which

    perimental results are not available (38). T

    challenges ofcijestimation notwithstanding (

    Box 2), one can specify interaction parame

    values, block sequences, and block lengths a

    then proceed to use large-cell SCFT simulatio

    to predict microphase structure candidates. If si

    ulations launched from a variety of random ini

    conditions consistently lead to a single defefree or defective, but still identifiable morpho

    gy, one can be reasonably confident that the stab

    structure has been identified. Defective morph

    ogies can also be converted to defect-free ones

    the deterministic seeding method illustrated

    Fig. 4. Conversely, if repeated large-cell SC

    simulations from uncorrelated random seeds tu

    up multiple structures differing in symmetry, th

    it is necessary to compare the free energies

    the competitive structures to establish which

    them is stable. For this purpose, a second ty

    of SCFT simulation is most efficienta unit-c

    simulation that attempts to converge a single u

    cell of a candidate structure within the confinof the symmetry constraints of that phase (4

    Such a simulation seeks a saddle-point condit

    on the Hamiltonian with respect to the comp

    field variables while simultaneously minimiz

    H(the mean-field free energy) with respect to t

    lattice parameters of the unit cell.

    State-of-the-art numerical methods for lar

    cell simulations are based on spectral collocat

    (or pseudospectral) techniques with plane wa

    bases (38, 42). Fast Fourier transforms are us

    to switch between real-space and reciproc

    space representations of the fields, allowing

    efficient evaluation of the operators, forces, a

    energies embodied in SCFT. Large 3D calcutions using up to 5123 = 1.34 108 plane wa

    or grid points, such as those shown in Fig

    require the use of parallel algorithms imp

    mented either on clusters of single or multico

    central processing units or, more recently, on

    single graphics processing unit containing

    to 500 lightcores. The method of choice

    unit-cell SCFT simulations is a Galerkin spec

    technique developed by Matsen and Schick t

    uses Fourier basis functions possessing the sy

    metry of the phase being considered (41). A

    though the method is not well suited to paral

    computing, a smaller number of symmetry-adap

    basis functions are generally required for ac

    rate unit-cell simulations than the plane wav

    used for spectral collocation in parallelepip

    cells (43).

    In spite of these advances in numerical SCF

    the tool is primarily used to predict self-assemb

    given a block polymer design. More useful fro

    the standpoint of applications is the inve

    problemnamely, the identification of polym

    designs that will self-assemble into a specif

    morphology. Though little progress has be

    made to date on this problem, techniques su

    as inverse Monte Carlo, in principle, could

    Fig. 4.Large-cell SCFT morphol-ogies obtained for an incompres-sible ABC triblock terpolymer melt

    using parameters (fA= 0.275,fB=0.550,fC= 0.175,cABN= 13, cACN= 35, cBCN = 13) and a periodicsimulation cell size commensuratewith 3-by-3-by-3 unit cells of theFddd(O70) phase. (A) A highly de-fective structure is obtained froma simulation initialized using anunstructured, random initial fieldconfiguration. (B) A defect-free mor-phology comprising 27 unit cellsof the O70 phase emerges from asimulation initialized using a smallnumber of field harmonics consist-ent with the Fddd space group.

    Both simulations were conducted onan NVIDEA Tesla C2070 graphicsprocessing unit.

    www.sciencemag.org SCIENCE VOL 336 27 APRIL 2012

    RE

  • 8/11/2019 P12_S_Bates_Multiblock Polymers - Panacea or Pandora's Box

    8/8

    mated to SCFT toward this end. It is also no-

    table that marked advances have transpired in

    algorithms for FTS simulations, which do not

    rely on the mean-field approximation (38, 44).

    These techniques can be used to explore a wide

    range of fluctuation-mediated phenomena and,

    when combined with thermodynamic integration

    and Gibbs ensemble methods, can yield accurate

    order-disorder phase boundaries for complex

    block polymers.

    Outlook

    Do more blocks presage a panacea or Pandoras

    box? At a minimum, thermoplastic elastomers

    requirek= 2 and n = 3 as evidenced by PS-PI-

    PS, introduced in the 1960s and still the most

    successful block copolymer product in the world-

    wide marketplace. Increasing n at fixed k = 2

    generates a host of new opportunities without

    conceptual difficulty (see Box 1) and at marginal

    or no added cost, provided that the appropriate

    synthetic tools are available. Recent advances

    in olefin polymerization chemistry, such as the

    dual-catalystmediated chain-shuttling mecha-

    nism (45), underscore this point. Increasing kquickly complicates the practical design of new

    materials yet challenges our imagination with

    unbounded opportunities. It is not unreason-

    able to claim that virtually any structure (domain

    morphologies and packing symmetry) can be

    created at length scales between roughly 5 nm

    and 1 mm using block polymers, while main-taining robust flexibility over individual block

    chemical and physical properties. Every (k,n)

    enumeration has the potential to produce a

    unique material. The resulting compounds may

    address engineering goals directly (e.g., multi-

    functional plastics) or enable a host of other

    products [for example, as intermediates in drugdelivery, as scaffolds that guide the assembly of

    inorganic hard materials (46), or for pattern for-

    mation in the manufacturing of microelectronics

    (3)]. Realizing these tantalizing prospects requires

    overcoming challenges posed by complexity, a

    dilemma that throttles many modern technolog-

    ical ambitions ranging from biological systems

    to global climate prediction. Here, the approach

    seems clear: integration of engineering goals, cre-

    ative chemistry, and predictive theoretical tools,

    augmented by continued advances in structural

    characterization methods. Fortunately, even mi-

    nor extrapolation in multiblock complexity (e.g.,

    k= 3,n = 4) offers a glimpse of what is possible

    in the future.

    References and Notes1. F. S. Bates, G. H. Fredrickson,Phys. Today52, 32 (1999).

    2. F. Meng, Z. Zhong, J. Feijen,Biomacromolecules10, 197

    (2009).

    3. R. Ruizet al., Science 321, 936 (2008).

    4. F. S. Bates, G. H. Fredrickson, D. Hucul, S. F. Hahn,

    AIChE J. 47, 762 (2001).

    5. T. J. Hermel, S. F. Hahn, K. A. Chaffin, W. W. Gerberich,

    F. S. Bates, Macromolecules 36 , 2190 (2003).

    6. W. Zheng, Z.-G. Wang,Macromolecules 28, 7215

    (1995).

    7. J. Chatterjee, S. Jain, F. S. Bates,Macromolecules 40,

    2882 (2007).

    8. C. Auschra, R. Stadler,Macromolecules26, 2171 (1993).

    9. Y. Mogi et al., Macromolecules 27, 6755 (1994).

    10. S. Christodoulou, P. Driva, H. Iatrou, N. Hadjichristidis,

    Macromolecules41, 2607 (2008).

    11. K. A. Dill, S. B. Ozkan, M. S. Shell, T. R. Weikl,Annu. Rev.

    Biophys.37, 289 (2008).

    12. M. Szwarc, M. Levy, R. Milkovich,J. Am. Chem. Soc. 78,

    2656 (1956).13. A. Hirao, T. Higashihara, M. Nagura, T. Sakurai,

    Macromolecules39, 6081 (2006).

    14. C. J. Hawker, K. L. Wooley,Science 309, 1200 (2005).

    15. K. Matyjaszewski,Science 333 , 1104 (2011).

    16. E. A. Jackson, M. A. Hillmyer,ACS Nano 4, 3548 (2010).

    17. N. A. Lynd, M. A. Hillmyer,Macromolecules 38, 8803

    (2005).

    18. J. M. Widin, A. K. Schmitt, A. L. Schmitt, K. Im,

    M. K. Mahanthappa, J. Am. Chem. Soc. 134 , 3834

    (2012).

    19. S. Guoet al., Chem. Mater. 18, 1719 (2006).

    20. A. O. Moughton, M. A. Hillmyer, T. P. Lodge,

    Macromolecules45, 2 (2012).

    21. A. J. Meuler, M. K. Mahanthappa, M. A. Hillmyer,

    F. S. Bates, Macromolecules 40 , 760 (2007).

    22. M. K. Kiesewetter, E. J. Shin, J. L. Hedrick,

    R. M. Waymouth, Macromolecules 43, 2093 (2010).

    23. K. Okamoto, C. K. Luscombe,Polymer Chem. 2, 2424

    (2011).

    24. V. Bellas, H. Iatrou, N. Hadjichristidis,Macromolecules

    33, 6993 (2000).

    25. E. Helfand,Macromolecules11, 682 (1978).

    26. L. Leibler,Macromolecules 13, 1602 (1980).

    27. E. W. Cochran, C. J. Garcia-Cervera, G. H. Fredrickson,

    Macromolecules39, 2449 (2006).

    28. J. Qin, F. S. Bates, D. C. Morse,Macromolecules43, 5128

    (2010).

    29. K. Hayashida, T. Dotera, A. Takano, Y. Matsushita,

    Phys. Rev. Lett.98, 195502 (2007).

    30. SCFT calculations support the two-domain and

    three-domain lamellar morphologies illustrated in

    Fig. 3A.

    31. M. J. Bluemle, J. Zhang, T. P. Lodge, F. S. Bates,

    Macromolecules43, 4449 (2010).

    32. P. M. Chaikin, T. C. Lubensky, Eds.,Principles of

    Condensed Matter Physics (Cambridge Univ. Press,

    Cambridge, 1995).

    33. T. Goldacker, V. Abetz, R. Stadler, I. Erukhimovich,

    L. Leibler, Nature 398 , 137 (1999).

    34. Q. Chenet al., Science 331 , 199 (2011).

    35. S. Lee, M. J. Bluemle, F. S. Bates,Science 330, 349

    (2010).

    36. J. Zhang, S. Sides, F. S. Bates,Macromolecules 45, 2

    (2012).

    37. S. F. Edwards, Proc. Phys. Soc. London 88, 265 (196

    38. G. H. Fredrickson, The Equilibrium Theory of

    Inhomogeneous Polymers (Oxford Univ. Press, Oxford

    2006).

    39. M. W. Matsen,J. Phys. Condens. Matter14, R21 (20

    40. G. H. Fredrickson, V. Ganesan, F. Drolet,Macromolec

    35, 16 (2002).

    41. M. W. Matsen, M. Schick,Phys. Rev. Lett.72, 2660 (19

    42. G. Tzeremes, K. O. Rasmussen, T. Lookman, A. Saxen

    Phys. Rev. E65, 041806 (2002).

    43. P. Stasiak, M. W. Matsen,Eur. Phys. J. E34, 110 (20

    44. E. M. Lennon, K. Katsov, G. H. Fredrickson, Phys. Re

    Lett.101 , 138302 (2008).

    45. D. J. A rriola, E. M. Carnahan, P. D. Hustad, R. L. KuhlmT. T. Wenzel, Science312, 714 (2006).

    46. M. Stefiket al., Chem. Mater. 21, 5466 (2009).

    47. E. Lucas,Thorie des Nombres (Gauthier-Villars, Pari

    1891).

    48. J. Riordan,An Introduction to Combinatorial Analysis

    (Princeton Univ. Press, Princeton, NJ, 1978).

    49. J. Kyte, R. F. Doolittle,J. Mol. Biol. 157, 105 (1982

    50. J. W. Ponder, F. M. Richards,J. Mol. Biol. 193, 775

    (1987).

    51. F. S. Bates, G. D. Wignall,Phys. Rev. Lett. 57, 1429

    (1986).

    52. T. P. Russell, R. P. Hjelm Jr., P. A. Seeger,

    Macromolecules23, 890 (1990).

    53. A. S. Zalusky, R. Olayo-Valles, J. H. Wolf, M. A. Hillm

    J. Am. Chem. Soc. 124 , 12761 (2002).

    Acknowledgments:This work was supported by the Mater

    Research Science and Engineering Centers Program of the

    NSF under award numbers DMR-0819885 (F.S.B., M.A.H., T.

    and DMR-1121053 (G.H.F., K.T.D.). C.M.B. thanks C. G. Willso

    for his support.

    Supplementary Materialswww.sciencemag.org/cgi/content/full/336/6080/434/DC1

    Supplementary Text

    10.1126/science.1215368

    REVIEW