08.06.2013 Views

Supramolecular Polymerizations

Supramolecular Polymerizations

Supramolecular Polymerizations

SHOW MORE
SHOW LESS

You also want an ePaper? Increase the reach of your titles

YUMPU automatically turns print PDFs into web optimized ePapers that Google loves.

Macromol. Rapid Commun. 2002, 23, 511–529 511<br />

Review: Unimers of both natural and synthetic origin<br />

self-assemble into linear, helical, columnar, planar and<br />

three-dimensional structures depending upon the functionality<br />

of supramolecular interactions. Recent reports<br />

describing the mechanism of formation, properties and<br />

possible applications of these systems are critically<br />

reviewed. The assembling of one-dimensional systems<br />

produces equilibrium polymers showing a length distribution<br />

and a degree of polymerization that may far exceed<br />

that of typical condensation polymers. Their growth may<br />

occur by a step-by-step process akin to polycondensation,<br />

and by cooperative processes such as helical growth or<br />

growth coupled to liquid crystallinity. Of particular interest<br />

are functional systems based on the coupling of a<br />

chemical reaction to supramolecular polymerization, and<br />

systems based on a covalent polymer hosted within the<br />

cavity of a supramolecular one. The assembly of two and<br />

three-dimensional systems occurs through a process akin<br />

to crystallization. The supramolecular organization of<br />

amphiphiles such as block copolymers is currently well<br />

described by the mean-field theory of unstable modes in<br />

homogeneous melts. An alternative, less sophisticated<br />

approach considers the growth of specifically designed<br />

building blocks. Possible applications are in areas that<br />

expand the uses of covalent polymers, electrochemical<br />

<strong>Supramolecular</strong> <strong>Polymerizations</strong><br />

Alberto Ciferri<br />

Dipartimento di Chimica e Chimica Industriale, Università di Genova, Via Dodecaneso 31, 16146 Genova, Italy<br />

E-mail: cifjepa@chimica.unige.it<br />

Keywords: assemblies; host-guest systems; hydrogen bonding; supramolecular structures;<br />

Contents<br />

1 Definitions. Classification of <strong>Supramolecular</strong> Polymers<br />

2 The Bond. Site and Shape Recognition<br />

3 Functionality of the Unimer and Dimensionality of<br />

the Assembly<br />

4 Theory of <strong>Supramolecular</strong> Polymerization<br />

4.1 Linear, Helical Assemblies<br />

4.2 Multidimensional Assemblies<br />

5 Polymers from <strong>Supramolecular</strong> Polymerization<br />

5.1 Linear Chains and Applications<br />

5.1.1 H-Bonded Polymers<br />

5.1.2 H-Bonded Networks<br />

5.1.3 Coordination Polymers<br />

5.1.4 Rheology and Polymer-Like Properties<br />

5.2 Helical Chains and Functional Systems<br />

and photonic devices, ion-selective channels, separation<br />

processes, microengines mimicking the performance of<br />

biological systems, storage of sequential information, biocompatible<br />

and patterned surfaces, sensors. A classification<br />

including additional systems that have been described<br />

as supramolecular polymers is presented.<br />

Disk-shaped, three blades molecule: formation of a helical<br />

assembly when the blades assume a propeller-like conformation.<br />

(i Am. Chem. Soc. 2000 and 2001.)<br />

5.3 Columnar and Micellar Assemblies<br />

5.3.1 Evidence for the SLC Mechanism<br />

5.3.2 Helical-Columnar Mechanism<br />

5.3.3 Supermolecules. Tubular Assemblies<br />

5.4 Host/Guest Polymeric Assemblies<br />

5.5 Planar Assemblies<br />

5.6 Composite and 3D Assemblies<br />

6 Topics for Further Investigation<br />

1 Definitions. Classification of <strong>Supramolecular</strong><br />

Polymers<br />

<strong>Supramolecular</strong> polymerization is the process of forming<br />

long sequences of bifunctional unimers linked only by<br />

main-chain noncovalent bonds. Theoretically well<br />

defined growth mechanisms assist the process, and extension<br />

of the underlying concepts to planar and three-<br />

Macromol. Rapid Commun. 2002, 23, No. 9 i WILEY-VCH Verlag GmbH, 69469 Weinheim 2002 1022-1336/2002/0906–0511$17.50+.50/0


512 A. Ciferri<br />

dimensional growth is possible. [1] While the assembly<br />

produced by supramolecular polymerization is a supramolecular<br />

polymer, [2] the reverse is not true. Several systems<br />

have been reported and defined as supramolecular<br />

polymers that do not conform to the growth mechanisms<br />

and theory of supramolecular polymerization.<br />

A supramolecular polymer (SP) can be defined broadly<br />

as a system characterized by non-bonded interactions<br />

among repeating units. Such a broad definition includes<br />

all systems that have been described as SPs although, due<br />

to its general nature, it does not suggest immediately the<br />

structural features or potential applications of this exciting<br />

class of new materials. In fact, even a molecular polymer,<br />

based on covalently bonded repeating units, displays<br />

high order structural organization controlled by supramolecular<br />

interactions. The occurrence of supramolecular<br />

interaction is so widespread that even an organic crystal<br />

is described as a supramolecular assembly. [3] Attempts to<br />

restrict the definition of SPs, for instance to systems displaying<br />

polymer-like properties in dilute solution, have<br />

been made. [4] The problem is not a simple one since several<br />

variables control the degree of supramolecular polymerization<br />

(DPÞ. Moreover, the term supramolecular has<br />

great appeal since it invites the use of unifying concepts<br />

that cut across the traditional boundaries between colloid,<br />

polymer and solid state science.<br />

Rather than attempting to further clarify the definition<br />

of SP, it is useful to present a classification of the various<br />

systems that have been reported. A possible classification,<br />

based on assembling mechanisms, is schematized in<br />

Figure 1. It offers a glimpse of the impressive growth that<br />

has occurred over the past decade and contextually highlights<br />

the SPs produced by supramolecular polymerization<br />

that form the core of the present review. The reference<br />

model of the classical covalent chain resulting from<br />

molecular polymerization of small bifunctional monomers<br />

is schematized at the top of Figure 1. The selfassembling<br />

chain is an open one, meaning that, in principle,<br />

it can grow to a distribution of large DPs, irreversible<br />

in solution and under a wide range of external variables.<br />

Figure 1. Classification of supramolecular polymers. Class A<br />

(reversible polymers obtained by supramolecular polymerization)<br />

is the main topic of this article.<br />

Class A. The major components of this class are equilibrium<br />

polymers based on processes that can appropriately<br />

be regarded as supramolecular polymerizations. [1, 2] The<br />

linear chains are self-assembled, open, growing to a distribution<br />

of DPs, and in a state of thermodynamic equilibrium<br />

sensitive to solvent type, concentration and external<br />

variables. The geometrical shapes in the scheme of<br />

class A (Figure 1) remind that unimers in supramolecular<br />

polymerization can be of several forms and sizes. In particular,<br />

the unimer can be a large supramolecular aggregate,<br />

a supermolecule, [2] a covalent polymer (e.g., a globular<br />

protein) in which case the SP will actually be a polymer<br />

of polymers. In class A we may also include SPs<br />

based on unimers with functionality A2, when a variety of<br />

multidimensional assemblies (helical, planar, 3D)<br />

becomes possible. Examples of linear systems are hydrogen-bonded<br />

polymers, [5–17] coordination polymers [18–20]<br />

and also micelles. [21–24] Examples of more complex geo-<br />

Alberto Ciferri is Chemistry Professor at the University of Genoa, as well as Visiting Professor<br />

at Duke University, Durham, North Carolina (1975-present). He has authored about<br />

200 original papers, books and patents mostly in the areas of rubber elasticity, biological<br />

and synthetic fibers, interactions between salts and macromolecules, liquid crystals, and<br />

supramolecular assemblies. He received his D.Sc. degree in physical chemistry from the<br />

University of Rome, and held positions as Scientist at Monsanto Co. and Director of<br />

Research at the National Research Council. He currently is President of the Jepa-Limmat<br />

Foundation, supporting advanced education in developing countries.


<strong>Supramolecular</strong> <strong>Polymerizations</strong> 513<br />

metries are helical, columnar, tubular soluble [25–44] or<br />

fibrous proteins, [45] S-layers, [46] composite systems such<br />

as block copolymers [47–49] and the tobacco mosaic virus<br />

(TMV). [50] Random networks and blends stabilized by<br />

multifunctional supramolecular linkages have also been<br />

reported. [51–55]<br />

Class B. This class includes self-assembled structures<br />

formed by supramolecular binding of monofunctional<br />

unimers. Such unimers cannot undergo open supramolecular<br />

polymerization, but can nevertheless form closed<br />

assemblies involving both low- and high-molecularweight<br />

species. Classical host/guest complexes, [56] base<br />

pairing of simple nucleoside [57] and supermolecules are<br />

low-molecular-weight examples. Polymeric examples<br />

described as SPs include side-chain binding of a monofunctional<br />

unimer to a covalent chain. For instance, Kato<br />

and FrØchet first reported the binding of the monofunctional<br />

mesogen stilbazole to the side chains of a nonmesogenic<br />

polymer functionalized with pendant benzoic<br />

acid groups. [58] Additional examples are double-, and triple-chain<br />

assemblies, and globular structures unable to<br />

grow further when complementary monofunctional sites<br />

are internally saturated.<br />

Class C. A number of SPs displaying novel supramolecular<br />

features were obtained by superimposition of covalent<br />

and supramolecular bonds. These systems are selfassembling<br />

but show irreversible DPs. The supramolecular<br />

organization may either precede, be simultaneous to,<br />

or follow the formation of covalent bonds. Examples of<br />

the first type include the rotaxane and catenane polymers<br />

described by Stoddart and coworkers, [59–61] the growth of<br />

dendrimers though successive generations, [62] and other<br />

attempts to stabilize a supramolecular assembly by the<br />

subsequent formation of covalent bonds. [63, 64] The final<br />

covalent system may retain specific supramolecular features,<br />

or the precursor supramolecular organization may<br />

just be a step of a supramolecularly assisted synthesis of<br />

a complex structure. Examples in which the supramolecular<br />

and the molecular order are simultaneously formed<br />

are the synthesis of dendrons possessing polymerizable<br />

functionality at their focal points, as reported by Percec<br />

and Schlüter. [65, 66] These assemblies display most interesting<br />

composite architectures such as columns of disks<br />

hosting the dendrons, with the main covalent chain running<br />

in the center of each column. [67] Cases in which the<br />

covalent structure occurs before the supramolecular one<br />

include the dendronization of a covalent polymer,<br />

reported for instance by Tomalia and coworkers, [68] and<br />

the self-assembled monolayers (SAMs) regarded as<br />

supramolecular assemblies of short hydrocarbon chains<br />

covalently grafted to a gold surface. [69]<br />

Class D. The class of engineered assemblies includes<br />

systems that do not form spontaneously ordered structures<br />

under normal conditions. Their ordered structurization<br />

is based on controlled methods of deposition or<br />

synthesis. Their classification as SPs can be justified<br />

since elements of supramolecular interaction still assist<br />

the final organization. Examples are the layered assembly<br />

of complementary polyelectrolytes obtained by step-wise<br />

deposition under kinetic control, [70] and polymer brushes<br />

prepared by grafting a polymer chain over a selfassembled<br />

monolayer of an initiator. [71] Both approaches<br />

allow a fine tuning of surface properties, complemented<br />

by patterning possibilities. Tailored performance in applications,<br />

such as biocompatibility, biocatalysis, integrated<br />

optics and electronics, are possible.<br />

[70, 71]<br />

The following sections detail concepts and results relevant<br />

to supramolecular polymerization and SPs in class A<br />

systems. In particular, Section 4 summarizes the theoretical<br />

framework of polymerization mechanisms [1] forming<br />

the basis for the critical analysis of the experimental data<br />

to be presented in Section 5. Some aspects already<br />

described in preceding reviews/analyses by the author [1]<br />

(focusing on mechanisms), by Lehn, [2] Zimmermann and<br />

coworkers, [5] Meijer and coworkers [4] (focusing on chemical<br />

features) are briefly summarized, placing more<br />

emphasis on recent data and concepts.<br />

2 The Bond. Site and Shape Recognition<br />

The cement holding well organized supramolecular structures<br />

requires the description of: (i) interaction between<br />

specific sites, (ii) site distribution and (iii) shape complementarity<br />

of the unimers (cf. ref. [1] for a detailed discussion).<br />

The relevant interactions are schematized in Figure<br />

Figure 2. Forces assisting supramolecular organization.


514 A. Ciferri<br />

Figure 3. (a) Bifunctional units with binding sites along the N<br />

and S (or E and W) directions, yielding linear polymers. (b) Tetrafunctional<br />

units with two sites along N and S, and two sites<br />

along N-E and S-E (same side of the assembly), yielding linear<br />

or helical double chains. (c) Tetrafunctional units with sites at<br />

right angles within the cross-sectional or equatorial area of the<br />

unimer, yielding planar polymers. (d) Hexafunctional units with<br />

sites as in (c), and two additional sites on the flat surfaces or<br />

poles, yielding three-dimensional polymers.<br />

2. Classical supramolecular interactions (Coulombic,<br />

hydrogen and van der Waals bonds) are localized at specific<br />

sites or atoms of the unimers. These sites may be<br />

distributed at discrete locations over the surface of the<br />

unimers: the direction of interaction determines the functionality<br />

of the unimer and, in turn, the dimensionality of<br />

the assembly (cf. next section and Figure 3). These localized<br />

interactions are described by the respective set of<br />

potential functions involving combinations of point<br />

charges, dipolar interaction and separation distances. Several<br />

combinations of the above interactions may occur<br />

over the surface of the unimer, additively contributing to<br />

the overall binding free energy. [56, 72] Cooperative effects<br />

(when the formation of the first pairwise interaction<br />

increases the binding constant at successive sites along<br />

the chain) are also possible. In the case of H-bonds that –<br />

due to their strong directionality – are a primary source<br />

of stabilization of several SPs in class A, the parallel or<br />

antiparallel arrangement of multiple bonds may increase<br />

or decrease the product of the single binding constants on<br />

account of secondary electrostatic interaction.<br />

[5, 6]<br />

In addition to the above site-localized classical interactions,<br />

other stabilizing interactions play an important role<br />

in polymeric assemblies. [1] The solvophobic bond is<br />

responsible for the micellization of amphiphiles in the<br />

presence of a solvent. Even in the absence of a solvent,<br />

the incompatibility of amphiphilic components produces<br />

their ordered microsegregation. These interactions can be<br />

described by thermodynamic parameters that control<br />

micro- and macrophase separations. For instance, the<br />

Flory-Huggins thermodynamic parameter v plays a primary<br />

role in the theoretical description of microsegregation<br />

in block copolymers (cf. ref. [1, 81] and Section 4.2).<br />

The occurrence of liquid crystallinity in solutions of<br />

several polymeric assemblies is an example of hierarchical<br />

structurization (“macroscopic expression of molecular<br />

recognition” [2] ). Again, a thermodynamic effect (e.g.,<br />

volume exclusion resulting from the shape anisotropy of<br />

rigid SPs) is the primary driving force for structurization.<br />

Note that it is convenient to distinguish the role of shape<br />

in the stabilization of individual unimers (shape I effect,<br />

cf. (iii) above) from the role of shape anisotropy in the<br />

development of liquid crystallinity by rigid SPs (shape II<br />

effect, cf. ref. [1] and Section 4.1).<br />

3 Functionality of the Unimer and<br />

Dimensionality of the Assembly<br />

The assessment of unimer functionality is a primary<br />

requirement for determining the dimensionality of the<br />

assembly. Bifunctional rod-like, disk-like or spherical<br />

unimers (see scheme in Figure 3a) having binding sites<br />

pointing toward the North and South directions (or<br />

toward East and West) yield linear polymers. Note that it<br />

is the directionality of the interaction that specifies the<br />

functionality, e.g., the same functionality is assumed if<br />

four H-bonds rather than a single one point in the same<br />

direction.<br />

Increasing the functionality of the unimers produces<br />

more complex structures. The presence of two additional<br />

sites produces extended or helical double chains, [1, 7] provided<br />

the additional sites are located at the same side of<br />

the assembly (e.g., NE and SE, Figure 3b). However, planar<br />

assemblies are expected when four sites point toward<br />

perpendicular directions within a cross-section of cylindrical<br />

and disk-like unimers, or the equatorial plane of a<br />

spherical unimer (Figure 3c). The occurrence of two<br />

additional sites on the flat surfaces of cylinders and disks,<br />

or the poles of a sphere, generates a three-dimensionally<br />

ordered network (Figure 3d).<br />

The symbols +, –, 0, and 9, in Figure 3 indicate the<br />

functionality and refer to any possible localized supramolecular<br />

bond (Coulombic, hydrogen, van der Waals). In<br />

the case of non-localized effects, such as incompatibility<br />

and shape II-induced mesophases, a more uniform distribution<br />

of repulsive interaction ought to be assumed. In a<br />

few cases a polymer is undoubtedly formed, but the specification<br />

of unimer size and shape may not be straightforward.


<strong>Supramolecular</strong> <strong>Polymerizations</strong> 515<br />

Figure 4. Theory of linear supramolecular polymerization.<br />

Schematic variation of the length (or DP) of a growing linear or<br />

helical assembly with the unimer concentration C according to<br />

three different growth mechanisms. MSOA: multi-stage open<br />

association, [1, 74] HG: helical growth, [26] SLC: open supramolecular<br />

liquid crystal. [1, 28– 30] C* is the critical concentration for helical<br />

growth, C i is the critical concentration for the formation of<br />

the mesophase. [77]<br />

4 Theory of <strong>Supramolecular</strong> Polymerization<br />

4.1 Linear, Helical Assemblies<br />

There are three different supramolecular polymerization<br />

mechanisms by which a linear or helical polymer could<br />

assemble consistently with the schemes in Figure 3a and<br />

b. [1]<br />

(i) Multistage open association (MSOA) resembles the<br />

step-by-step mechanism of the molecular polycondensation<br />

of bifunctional unimers. [73] The increase in the degree<br />

of polymerization of the growing assembly with the (total)<br />

[1, 74, 75]<br />

unimer concentration C is given by<br />

C L [M0/(4KNa)] N [(DP w) 2 –1] (1)<br />

where M0 is the unimer molar mass, Na the Avogadro’s<br />

number and K is the site binding constant. A plot of<br />

Equation (1) is schematized in Figure 4a. A continuous<br />

increase in DP, or length L of the assembly, with concentration<br />

is expected and the rate of increase is strongly<br />

affected by the binding constant. For K a 10 6 m –1 , corresponding<br />

to one or two H-bonds per site, only oligomers<br />

(DP a 10) occur in dilute (a1%) solution. However, in<br />

the case of the ureidopyrimidone polymers reported by<br />

Meijer and coworkers, [4, 52] characterized by four H-bonds<br />

per site and K A 10 7 , extremely high DPs (e 1000) may<br />

be reached in dilute isotropic solution. The familiar Car-<br />

others equation and the control of DP by monofunctional<br />

unimers [73] apply to MSOA.<br />

(ii) Helical growth (HG) is achieved when step-by-step<br />

growth is reinforced by an intra-assembly cooperative<br />

effect, [26] originating from a peculiar functionality such as<br />

that schematized in Figure 3b. Since more bonds per<br />

unimers occur with respect to the situation in Figure 3a,<br />

Kh A K and a critical unimer concentration C*occurs at<br />

which the MSOA regime encroaches helix formation<br />

with a sudden increase in DP according to [26]<br />

DPn =(Ch/C*) 1/2 N r –1/2 (2)<br />

where r s 1 is the familiar cooperativity parameter<br />

(easily in the order of 10 –8 ) and Ch is the concentration of<br />

helical polymer (increasing when C A C*). Figure 4b<br />

shows the step-jump increase in DP, orL, accompanying<br />

the nucleation of the helix at C*. The theory was intended<br />

to describe the reversible aggregation of globular proteins<br />

attaining extremely high DPs(A4000, cf. Section 5.2).<br />

Van der Schoot and coworkers [76] have recently generalized<br />

Oosawa’s theory to cover general situations with<br />

Kh A K, including high and weakly cooperative helical<br />

aggregation. The theory was applied to the case of columnar<br />

assemblies of chiral discotic molecules (cf. Section<br />

6.3.2).<br />

(iii) Growth-coupled-to-orientation (SLC) is alternatively<br />

described as the open supramolecular liquid crystal.<br />

[77] It is an inter-assembly cooperative process that<br />

encroaches MSOA causing a sudden enhancement of the<br />

step-by-step growth of bifunctional unimers at a critical<br />

concentration C i at which nematic order appears.<br />

[1, 21–23]<br />

The polymer must have considerable chain rigidity, [22] as<br />

expressed by its persistent (q) or deflection length (k),<br />

and the DP attained at C i may be approximated as<br />

DP V q/L0<br />

where L0 is the length of the unimer. The schematization<br />

of the theoretical behavior in Figure 4c shows the sudden<br />

jump of DP, orL, atC i (note C i is generally S C*), followed<br />

by a minor increase in L upon further increasing<br />

the unimer concentration. [23] Values of q for SPs frequently<br />

exceed the lm range, [1] enabling DPs higher than<br />

1000. Note that while attractive interactions stabilize the<br />

assembly along the N-S direction, no lateral (soft) anisotropic<br />

attraction among the formed assemblies is postulated<br />

to occur in the nematic state: only excluded volume<br />

(hard) interaction stabilizes the mesophase. In view of its<br />

fundamental significance, the difference between an open<br />

SLC and a conventional, molecular liquid crystal should<br />

be emphasized. In the latter, no association/dissociation<br />

equilibria accompany the transition from the isotropic<br />

solution: the nematic field has only an orienting effect. In<br />

the open SLC, development of orientation is simultaneous<br />

to an enhancement of supramolecular polymeriza-<br />

(3)


516 A. Ciferri<br />

tion. [77] Growth-coupled-to-orientation could in principle<br />

occur even for systems displaying only soft interactions,<br />

when mesophases are expected in the melt or in very concentrated<br />

solutions. [11] However, if growth due to alternative<br />

mechanisms (MSOA or HG) has produced a wormlike<br />

chain exceeding the persistent length (L A q) at these<br />

high concentrations, further growth by the SLC mechanism<br />

will not be relevant. The original theory, [20, 21] developed<br />

to explain the linear assembly of cylindrical<br />

micelles in nematic solutions, was later extended to discotic<br />

molecules in a wide range of concentrations at<br />

which hexagonal and higher-order phases were pre-<br />

dicted.<br />

[31, 78, 79]<br />

A common feature of the cases considered above is the<br />

occurrence of polydispersity. [80] It is also apparent that<br />

SPs attain average DPs that are concentration-dependent,<br />

but may nevertheless far exceed those obtained in molecular<br />

polycondensation (cf. ref. [73] and Section 5.1.4).<br />

4.2 Multidimensional Assemblies<br />

The above description of supramolecular polymerization<br />

for linear systems has involved the identification of<br />

proper unimers and corresponding growth mechanism.<br />

Unimers with functionality A2 can express strong supramolecular<br />

interactions in two and three dimensions. How<br />

can the approach developed for linear SPs be extended to<br />

multidimensional systems? General thermodynamic considerations<br />

regarding the growth of multidimensional<br />

assemblies were summarized by Israelachvili. [80] The<br />

standard chemical potential per unit (ln 0 ) decreases with<br />

the number n of aggregating units according to<br />

l 0 n = l 0 v + a kT/n p (4)<br />

where l 0 v is the bulk free energy for an infinite aggregate,<br />

a reflects the strength of contact energy, and p is a dimensionality<br />

index (p = 1 for linear systems, 1/2 for discs, 1/3<br />

for spheres). Elaboration of the approach leads to the<br />

expectation that, whenever p a 1, macroscopic aggregates<br />

(n ev) grow abruptly above a critical concentration by<br />

a mechanism corresponding to a crystallization. Thus, at<br />

variance with the linear systems, no concentrationdependent<br />

broad size distributions are expected and there<br />

is no need to invoke a growth-coupled-to-orientation<br />

mechanism. These considerations apply to the growth of<br />

monolayers, single lamella, and also to the growth of<br />

three-dimensional assemblies of unimers having symmetrical<br />

and equivalent functionality. For some 2D systems,<br />

finite-size effects may frustrate growth to infinite assemblies.<br />

[81] The description of the growth of 3D systems is<br />

also more complex whenever a preferential growth direction<br />

occurs, reflecting, for instance, the geometrical anisotropy<br />

of the polymer, or the presence of two components.<br />

In such cases it might be possible to follow the<br />

Figure 5. Schematic assembly of the side and top views of a<br />

rigid polyanion and a cationic surfactant. This assembly may<br />

grow longitudinally by stacking, and laterally by interdigitation<br />

and hexagonal packing taken (taken from ref. [81] ).<br />

modes of growth along the longitudinal and transversal<br />

directions.<br />

The situations that may be encountered are illustrated<br />

by the assembly of a rigid polyanion (e.g., DNA) and a<br />

cationic surfactant in water (Figure 5). [82] The observation<br />

of the conventional nematic phase expected for the DNA/<br />

H2O system is precluded by the strong reduction of solubility<br />

when the surfactant is present. Growth along the<br />

lateral direction (D) occurs by consecutive interdigitation<br />

of DNA helices with bound surfactant and produces<br />

aggregates with hexagonal symmetry. Growth along the<br />

longitudinal direction (L) may also occur due to the<br />

hydrophobic interaction occurring at the exposed North<br />

and South surfaces of the assembly.<br />

Provided L prevails over D, the nematic and hexagonal<br />

phases, crucial for the spatial orientation of the growing<br />

columns, should develop through the linear growthcoupled-to-orientation<br />

mechanism. On the other hand,<br />

when D prevails over L, planar growth leads to aggregates<br />

of infinite size (crystallization). Due to the fact that<br />

the helix and surfactant molecules are not chemically<br />

bound, it ought to be possible, through compositional<br />

control, to monitor growth along the two directions, possibly<br />

evidencing a liquid-crystalline phase, enhancing<br />

growth along the longitudinal direction. In general, one<br />

would expect a difference in the growth rate along the<br />

longitudinal and lateral directions. Indeed, such differences<br />

have been observed in block copolymers, when the<br />

components are chemically bound. [83]<br />

The foregoing considerations justify the empirical<br />

identification of linear growth components even in bulk<br />

composite assemblies, to be discussed in more detail in<br />

Section 5.6. A sounder mean-field approach was developed<br />

to interpret the solid-state morphology of (A)n–(B)m<br />

amorphous diblock copolymers exhibiting a microsegregation<br />

of components into cylindrical, lamellar or spherical<br />

domains. [48] The approach is based on the thermodynamic<br />

incompatibility of copolymer components that cannot<br />

be crystallized, theoretically shown to generate an<br />

unstable mode in an homogeneous, undiluted melt. [84–87]<br />

The formation of interfaces reduces the enthalpic cost of<br />

mixing (measured by the v parameter) but entails an<br />

entropy loss due to chain stretching to fill space uniformly.<br />

The latter depends upon the relative length and


<strong>Supramolecular</strong> <strong>Polymerizations</strong> 517<br />

Figure 6. Mean-field phase diagram for amorphous diblock<br />

copolymers (A)n-(B)m. Main phases are: lamellar (L), hexagonal<br />

cylindrical (H), closed packed spheres (CPS), disordered (DIS).<br />

f is the volume fraction of A segments. (i Am. Chem. Soc.<br />

1996 [87] ).<br />

flexibility of the A and B blocks. On this basis, cubic,<br />

hexagonal, lamellar, and other phases are predicted in<br />

discrete regions of v N N versus m/n phase diagrams as<br />

that illustrated in Figure 6 (N: total number of A and B<br />

units). A review of the most recent elaboration that unifies<br />

the weak [85] and strong [86] segregation regimes is<br />

given in the literature. [87] Noolandi et al. [88] extended the<br />

(self-consistent) mean-field approach to the calculation of<br />

the relative stability of liquid-crystalline phases occurring<br />

in solutions of copolymers composed of three flexible<br />

blocks.<br />

5 Polymers from <strong>Supramolecular</strong><br />

Polymerization<br />

5.1 Linear Chains and Applications<br />

5.1.1 H-Bonded Polymers<br />

This class of systems has attracted considerable interest<br />

due to the intrinsic characteristics of H-bonds, such as<br />

directionality and the possibility of increasing bond<br />

strength by multiple pairwise interactions. Figure 7 illustrates<br />

typical cases in which flexible or rigid segments<br />

are functionalized with groups able to form single or multiple<br />

H-bonds. The binding constant K is the primary<br />

parameter controlling DP in terms of the simple MSOA<br />

mechanism schematized in Figure 4a, predicting, according<br />

to Equation (1), K A 106 –107 m –1 needed for DP above<br />

the oligomeric range. If one takes K = 500 m –1 as an average<br />

value for a single H-bond (reported for the pyridine/<br />

benzoic acid dimerization), [7] it appears that at least 4 Hbonds<br />

are needed to produce DPs of interest. In fact, the<br />

polymer in Figure 7c based on the dimerization of ureidopyrimidone<br />

residues (K =66107 m –1 in CDCl3), [4] was<br />

Figure 7. <strong>Supramolecular</strong> polymers stabilized by main-chain<br />

links based on (a) one, [11] (b) three, [2, 15] and (c) four [4, 52] Hbonds.<br />

reported to attain DP in the order of 1000 in dilute, iso-<br />

[4, 52]<br />

tropic solution.<br />

In the case of polymer (b), based on the 3 H-bond<br />

scheme of rigid anthracene segments terminated by uracil<br />

(A-A) or pyridine residues (B-B), one therefore would<br />

not expect a significant DP to occur in virtue of the<br />

MSOA mechanism. In fact, no evidence for appreciable<br />

DP was reported for isotropic solutions. However, the<br />

development of liquid crystallinity evidenced the formation<br />

of large DPs in moderately concentrated solutions in<br />

organic solvents, [14] likely triggered by the mechanism of<br />

growth-coupled-to orientation.<br />

When the rigid segments in (b) were replaced by flexible,<br />

tartaric acid spacers, the occurrence of liquid crystallinity<br />

was observed only in undiluted (thermotropic) systems.<br />

[13] Polymer (c), based on flexible segments terminated<br />

by diacid (A-A) and dipyridil (B-B) residues forming<br />

single H-bonds, [8–12] displayed analogously only thermotropic<br />

behavior. The occurrence of liquid crystallinity<br />

in the melt does not provide evidence of significant polymerization.<br />

[1] In fact, the role of the growth-coupled-toorientation<br />

mechanism for worm-like chains displaying<br />

soft interaction was shown to be extremely small. [10, 11] On<br />

the other hand, Equation (1) offers compelling evidence<br />

that no large DPs are produced in the case of a single Hbond.<br />

Other linear H-bonded systems of interest include the<br />

polycaps [89] based on the polymerization of capsular host<br />

complexes (calixarenes) functionalized with urea. The<br />

formation of main-chain bonds occurred only when a<br />

guest molecule was hosted in the capsule (CHCl3 acted<br />

both as a capsule host and as a solvent). Whitesides and<br />

coworkers [16] who had discussed the formation of closed<br />

structures by the self-assembly of melanine and cyanuric<br />

acid earlier, reported the formation of linear nanorods by<br />

introducing a mismatch in the spacers of the AA and BB<br />

groups of these monomers. Ladder-type supramolecular<br />

polymers, [10] and ribbon-type polymers have also been<br />

reported. [17]


518 A. Ciferri<br />

Figure 8. (a) Subunits used for supramolecular networks. [10] (b)<br />

PEO/PPO block copolymer networks based on supramolecular<br />

(upper) and covalent crosslinks. [52] (c) Hydrogen-bond-induced<br />

compatibilization in a polymer blend. [53]<br />

5.1.2 H-Bonded Random Networks<br />

In the scheme of Figure 7 all unimers exhibit a complete<br />

match of the donor/acceptor components of either single<br />

or multiple H-bond units. If this match does not occur, or<br />

tetra and bifunctional unimers are mixed, planar or threedimensional<br />

networks are possible. [2] Networks based on<br />

triacids and bipyridine derivatives, or on tetrafunctionalized<br />

pyridine and difunctional benzoic acid compounds<br />

(cf. Figure 8a), have been reported. [10, 51] Single H-bonds<br />

connect chain segments emanating from tetrafunctional<br />

crosslinkages. Meijer et al. [52] have reported functionalized<br />

copolymers of propylene oxide and ethylene oxide<br />

exhibiting a strong four H-bond scheme (Figure 8b).<br />

These networks exhibit peculiar rheological features to<br />

be described below. H-bonding (Figure 8c) between<br />

polymers, such as poly(4-vinylpyridine) and poly(4hydroxystyrene),<br />

[53] was described as a factor promoting<br />

[55, 90]<br />

compatibility in polymer blends.<br />

5.1.3 Coordination Polymers<br />

The scheme of linear coordination polymerization was<br />

discussed by Lehn. [2] The unimers are ditopic ligands<br />

with two binding groups forming main-chain bonds<br />

through metal-ion coordination (Figure 9a). Several metal<br />

binding groups (bidentate, tridentate) and metal ions with<br />

tetra-, penta- and hexa-coordination are available. Among<br />

Figure 9. (a) Schematization of a linear coordination SP showing<br />

bidentate and tridentate metal binding group and metal ions<br />

with tetra-, penta-, and hexa-coordination. [2] (b) Degree of polymerization<br />

of Be(Bu2PO2)2 vs concentration in CHCl3 at room<br />

temperature. The chain backbone is schematized on the right. [18]<br />

(c) Self-assembly of a cobalt porphyrin polymer by coordination<br />

of two covalently attached pyridine ligands. [20]<br />

the earliest reports of soluble, reversible coordination<br />

polymers we find systems based on three-atom-bridging<br />

phosphinate groups connected by tetrahedral metal atoms<br />

reported by Ripamonti and coworkers [18] in 1968. Figure<br />

9b illustrates the variation of DP (by means of vapor<br />

pressure osmometry) with the concentration of beryllium<br />

dibutylphosphinate (Be(Bu2PO2)2) dissolved in CHCl3, a<br />

non-coordinating solvent. Fiber-forming properties, suggesting<br />

larger DPs, were exhibited by anisotropic gels<br />

occurring in more concentrated solution. Substantial evidence<br />

of depolymerization with dilution was observed,<br />

confirming the dynamic reversibility typical of supramolecular<br />

polymers. [4] The chain structure, deduced from Xray<br />

diffraction, is based on the alternate singly and triply<br />

bridged structure shown in Figure 9b.<br />

Among the most recent reports, [20] the functionalized<br />

porphyrin polymer in Figure 9c was shown to attain DP<br />

L 100 in a 7610 –3 m solution in CHCl3 (by means of<br />

size-exclusion chromatography (SEC)). Here, coordination<br />

occurs between the Co atom (hexa-coordination) and<br />

the two pyridine ligands.


<strong>Supramolecular</strong> <strong>Polymerizations</strong> 519<br />

Figure 10. Complex viscosity (Pa N s) vs frequency x of polymer<br />

C in Figure 4 between 30 and 558C, and of the equivalent<br />

polymer (bottom, 308C) exhibiting a 2 H-bond (taken from<br />

ref. [52] ).<br />

5.1.4 Rheology and Polymer-Like Properties<br />

It is certainly surprising to observe that SPs based on<br />

bonds that are weaker than covalent ones can nevertheless<br />

attain DP larger than obtained, for instance, from<br />

conventional polycondensation. As indicated above, DP<br />

L 1000 can be expected for polymer (c) in Figure 7 under<br />

thermodynamic equilibrium, whereas DP L 100 requires<br />

the use of irreversible conditions in the case of aliphatic<br />

polyamides. [73] Thus, SPs may exhibit strong growth and<br />

labile bonds. Which applications can be conceived for<br />

such systems? The spontaneous thermodynamic assembly<br />

q disassembly process allows variations of DP in<br />

response to temperature, concentration, and other external<br />

variables. Moreover, the concomitant readjustment of<br />

donor/acceptor partners even under constant values of<br />

these variables renders the SPs truly adaptive, self-healing,<br />

combinatorial materials. [2] It is important to distinguish<br />

cases in which the growth of the SP is controlled by<br />

a non-cooperative mechanism (when changes in the<br />

above variables produce relatively minor DP changes, cf.<br />

Figure 4a) from cases in which cooperative effects are<br />

operative (cf. Figure 4b or c). Materials exhibiting minor<br />

DP alterations under the influence of an external variable,<br />

while still allowing the persistence of appreciable DPs,<br />

offer opportunities in areas of conventional polymers. For<br />

instance, the beneficial value of a relatively large DP on<br />

the mechanical properties will not necessarily be accompanied<br />

by a prohibitively large melt viscosity under processing<br />

conditions, as is the case for covalent polymers<br />

(properties of materials undergoing major changes in DP<br />

due to changes in external variables will be considered in<br />

the following section).<br />

The expectations described above are fully supported<br />

by rheological studies on some of the H-bond systems<br />

described above. Figure 10 illustrates the viscoelastic<br />

Figure 11. Isothermal viscosity and normal forces vs shear rate<br />

for a solution of polycapsules (C L 3% in o-dichlorobenzene)<br />

(taken from ref. [89] ).<br />

Table 1. Applications of linear polymers.<br />

STRONG T-DEPENDENT RHEOLOGY AT LARGE DP: easy<br />

to flow, stronger in use<br />

EXTENDING DP of covalent polymers<br />

RECYCLING: with complete regeneration properties<br />

TUNABLE, SMART MATERIALS: adjusting properties to<br />

environmental variables (switches, etc.)<br />

DIFFERENT CORES: mechanical, conductivity, light emitting,<br />

catalytic properties<br />

SELF-REPAIRING: any structural damage<br />

STRUCTURAL CONTROL: in copolymers, high selectivity,<br />

alternation, chirality<br />

SUPRA e covalent<br />

behavior exhibited by the polymer in Figure 4c under<br />

small oscillatory deformation at various temperatures (M — n<br />

=8610 3 ). [52] The zero-shear viscosity at 308C is comparable<br />

to that shown by an unfunctionalized polydimethylsiloxane<br />

with M — n = 3610 5 , and appears to be<br />

1000 times larger than for the compound based on similar<br />

unimers linked by a 2 H-bond scheme. The strong non-<br />

Newtonian behavior reflects the interplay of polymer viscoelasticity<br />

and chain dissociation at higher temperature.<br />

At low frequency (x) and high temperature (T), the loss<br />

modulus is larger than the storage modulus, while the<br />

reverse was observed at higher x and lower T. Consistent<br />

data was exhibited by the reversible networks displayed<br />

in Figure 8b showing a plateau modulus (5610 5 Pa) six<br />

times larger than that for a corresponding covalent copolymer.<br />

Figure 11 is even a more stringent demonstration<br />

of persistent polymeric behavior in spite of reversible<br />

polymerization. [89] Normal forces attaining values in the<br />

order of 1000 Pa are indisputable evidence of polymeric<br />

behavior, and all data was reversible upon reducing the<br />

shear rate. The rheological behavior of a coordination<br />

polymer (Cu(II) tetraoctanoate in decalin) was inter-


520 A. Ciferri<br />

preted [90] in terms of a theory describing the chain-extending<br />

role of labile crosslinkages. [91]<br />

Applications. Applications of SPs have been suggested<br />

in areas that expand the applicability of covalent polymers<br />

(Table 1). The advantage of an easier processing in<br />

spite of a large DP has been commented above. Particularly<br />

significant is the possibility of increasing the DP of<br />

conventional polymers by main-chain supramolecular<br />

bonds. Under study is the extension of the DP of aromatic<br />

polyamides that are usually produced in the desirable<br />

high-molecular-weight range only by cumbersome syntheses.<br />

[92] Polymers based on long covalent segments<br />

extended by supramolecular bonds should be recyclable<br />

with complete recovery of properties. [92] The subsequent<br />

transformation of supramolecular bonds into covalent<br />

ones should allow the stabilization of large DPs and<br />

structures that would have been extremely difficult to<br />

synthesize directly. [38, 93] The thermodynamic control of<br />

DP and of the topology of networks should allow applications<br />

as tunable, smart materials responding to changes in<br />

variables, such as temperature, stress, and solvent permeation.<br />

[2, 52] By using different covalent segments in A-<br />

A/B-B unimers, the tunability could be expressed in<br />

mechanical, conductivity, light emitting, and catalytic<br />

properties. If changes in the above variables were occurring<br />

accidentally, the SP would have the capability of<br />

self-repairing. The high selectivity of molecular recognition<br />

should allow the selection of proper sequences in<br />

mixtures of more than two complementary unimers.<br />

5.2 Helical Chains and Functional Systems<br />

The original Oosawa theory (cf. Section 4.2) is based on<br />

unimers exhibiting the site distribution shown in Figure<br />

3b. Cooperation arises because each unimer makes two<br />

bonds along the linear sequence and two weaker ones<br />

along the helical pattern. The unimers should be large<br />

enough to avoid possible mismatches in the pattern.<br />

There has been no verification of the model using synthetic<br />

SPs. The model was elaborated to describe the formation<br />

of helices and microtubules (G q F transformation)<br />

by biological SPs. [1, 26] However, the verification of<br />

the helical growth model has been problematic even in<br />

the relatively simple case of actin, due to the simultaneous<br />

dephosphorylation of ATP normally bound to the<br />

protein. A study in which growth could be observed without<br />

the complication of ATP e ADP hydrolysis was performed<br />

by Korn on actin-ADP. [27] Actin polymerization<br />

occurs by increasing the ionic strength or temperature in<br />

isotropic solution at unimer concentrations below<br />

0.04 mg/ml. Filaments exceeding 11 lm, corresponding<br />

to a DP larger than 4000, are attained (higher values<br />

were observed with tubulin). [1, 28] The length distribution<br />

conforms to the most probable distribution and chain<br />

stoppers reduce DP, in line with theory. The diagram in<br />

Figure 12. (a) Concentration of helical (0) and oligomeric<br />

(6) actin vs unimer concentration showing the critical concentration<br />

as predicted by Oosawa’s theory (taken from ref. [29] ). (b)<br />

Schematization of translational movement resulting from the<br />

directional growth of F-actin (taken from ref. [25] ).<br />

Figure 12a exhibits the theoretically predicted occurrence<br />

of a critical unimer concentration at which helical growth<br />

begins, and the concentration of unimers and oligomers<br />

attains a constant value. [29] The recent generalization [76] of<br />

Oosawa’s theory to growth processes enhanced by cooperative<br />

effects has not yet been applied to specific chainlike<br />

sequences. It has however been applied to the growth<br />

of columnar systems to be discussed in the following section.<br />

The overall functioning of actin or tubulin as in vivo<br />

systems invites to consider the way in which the helical<br />

growth process is coupled to the dephosphorylation reaction.<br />

This coupling produces a dynamic function of the<br />

polymer [25, 30] that assists in processes, such as motility,<br />

contraction, and cell division. Figure 12b illustrates the<br />

processes believed to occur during the polymerization of<br />

actin filaments. The critical concentration of G-actin<br />

unimers is defined by the condition of equality between<br />

the sum of the assembly and disassembly rates at the two<br />

ends of the filament (the barbed and the pointed ends).<br />

Under ATP hydrolysis, the depolymerization at one end<br />

may be faster than polymerization at the other end. Thus,<br />

a cycling of actin monomers from one end to the other<br />

occurs during the growth of F-filaments, resulting in a<br />

translation of the polymer (tread-milling effect). A<br />

related dynamic instability effect controls the assembly q<br />

disassembly process of tubulin into microtubules. [25]<br />

Applications. The design principle of functional protein<br />

systems is based on the coupling of a supramolecular<br />

polymerization process to a chemical reaction.<br />

[25, 30]<br />

Understanding and possibly reproducing coupled<br />

mechanisms is the challenging road to microengines and<br />

other functional SPs mimicking mechanical properties of<br />

biological systems and engineered for new practical<br />

applications.


<strong>Supramolecular</strong> <strong>Polymerizations</strong> 521<br />

Figure 13. (a) Schematization of the assembly of surfactant<br />

and growth of micellar polymers. (b) Experimental data and theoretical<br />

phase diagram for the discotic amphiphile 2,3,6,7,10,11hexa(1,4,7-trioxoacetyltriphenylene)<br />

in D2O (taken from ref. [31] ).<br />

5.3 Columnar and Micellar Assemblies<br />

5.3.1 Evidence for the SLC Mechanism<br />

In this section we will consider amphiphilic unimers that<br />

unmistakably exhibit nematic phases appearing simultaneously<br />

with the onset of extensive polymerization, thus<br />

revealing a superimposition of the MSOA and SLC<br />

mechanisms. If the shape of the unimer is cylindrical or<br />

disk-like, the resulting SPs will be linear or discotic. In<br />

both cases, theory suggests that strong contact forces and<br />

rigidity along the longitudinal direction are involved (cf.<br />

Section 4.1.3).<br />

Odijk [22, 94] has discussed the experimental verification<br />

of the growth-coupled-to-orientation theory in the case of<br />

conventional spherical micelles that exhibit, upon<br />

increasing the surfactant concentration, the assembling<br />

sequence: dispersed molecules e spherical micelles e<br />

cylindrical (end-capped) micelles e linear polymer, the<br />

growth of linear polymers being associated to the transformation<br />

isotropic e nematic.<br />

The broad features of the experimental phase diagram<br />

were found to be in line with theory, although some discrepancies<br />

remain. [95, 96] Persistence length data confirmed<br />

the rigidity of the micellar polymer, but exhibited consid-<br />

erable scattering with q ranging from L0.02 to 10 lm; a<br />

likely value of L1 lm corresponds to a DP in the order of<br />

20000. There is no data on the formation of the nematic<br />

phase and on persistence length in the case of block copolymer<br />

micelles. Higher-order phases (hexagonal, lamellar,<br />

etc.), however, have been observed. [97, 98] The absence<br />

of a nematic phase was also noticed for several conventional<br />

surfactants. The growth-coupled-to-orientation theory<br />

predicts that the nematic phase can be skipped and<br />

only a direct isotropic e hexagonal columnar phase is<br />

observed for suitable combinations of contact forces and<br />

persistence length. [79] It is also expected that volume<br />

excluded effects [23] and a reduced chain stretching [99]<br />

cause some micellar growth even in isotropic solutions of<br />

block copolymers.<br />

A most detailed evidence for growth-coupled-to-orientation<br />

is provided by discotic molecules: ditopic structures<br />

possessing a disk-shaped core from which a number<br />

of flexible alkyl chains emanate. Molecularly dispersed<br />

disks would be expected to form conventional liquid crystals<br />

in virtue of their large excluded volume. Moreover,<br />

disks are able to aggregate into soluble columns due to a<br />

p-p stacking of the cores and solvophobic interaction of<br />

the side chains. As in other supramolecular polymerizations,<br />

columns may assemble due to the small equilibrium<br />

constant (MSOA), and growth can be reinforced by the<br />

occurrence of cooperative effects. Figure 13b displays<br />

the phase diagram of the discotic amphiphile<br />

2,3,6,7,10,11-hexa(1,4,7-trioxoacetyltriphenylene) in<br />

D20. [31] The diagram represents a match between experimental<br />

data and theoretical lines calculated according to<br />

the theory of growth-coupled-to-orientation, and includes<br />

the prediction of higher-order phases. [68] Growth is seen<br />

to occur simultaneously with the appearance of the<br />

nematic phase at a concentration of L20% at room temperature.<br />

It appears that, for this system, the balance of<br />

contact forces and flexural rigidity favors the occurrence<br />

of the nematic phase. The formation of the hexagonal<br />

phase observed at higher concentration is attributed to an<br />

improved packing efficiency with respect to the nematic<br />

phase. The diagram reveals a hierarchical evolution of<br />

the assembling process through higher-order phases:<br />

disks (I) e columns (N) e hexagonal columnar (H) e<br />

solid.<br />

5.3.2 Helical-Columnar Growth Mechanism<br />

Meijer and coworkers [32, 33] have synthesized a series of<br />

most interesting C3-symmetrical disk-like molecules that<br />

assembles into cylindrical stacks in virtue of both hydrogen<br />

and arene-arene interactions. The molecules shown<br />

in Figure 14a have large aromatic cores, H-bonding<br />

groups, and either achiral or chiral side chains. The latter<br />

varied in their polar character allowing the study of<br />

aggregation in either polar or nonpolar solvents. The for-


522 A. Ciferri<br />

Figure 14. (a) Disk-shaped, three blades molecule prepared by<br />

Meijer and coworkers [4] with side chains having polar, nonpolar,<br />

chiral, achiral character. [32, 33] (b) Formation of a helical assembly<br />

when the blades assume a propeller-like conformation. [100]<br />

(c) Schematic representation of the transition from dispersed<br />

disks to partially ordered and to fully ordered columns. [76]<br />

(i Am. Chem. Soc. 2000 and 2001.)<br />

mation of long columns was shown to occur in very dilute<br />

solution in hexane (10 –6 m), and a large association constant<br />

(10 8 m –1 ) was reported. [32] It was suggested that such<br />

an aggregation reflects the formation of helical columns<br />

through a cooperative process attributed to a conformational<br />

transition from flat to propeller shape for the blades<br />

of each disk, resulting in a maximization of interaction<br />

for the chiral helical assembly (Figure 14b). [100] Experimental<br />

data for the more polar molecules in dilute butanol<br />

(2.4610 –4 m) [33] revealed a sequence of two association<br />

steps upon temperature changes. The postulated process<br />

is schematized in Figure 14c: starting with an isotropic<br />

dispersion of disks, a decrease in the temperature causes<br />

the formation of low-DP achiral aggregates stabilized by<br />

non-cooperative interaction, followed at a lower critical<br />

temperature by the cooperative formation of helical columns<br />

with DP attaining the 1000 range. [76]<br />

The theoretical description of the processes described<br />

above was formulated by van der Schoot and coworkers. [76]<br />

The occurrence of the two regimes was described in terms<br />

of the cooperativity parameter r. The situation r =1is<br />

equivalent to the binding of unimers into disordered aggregates<br />

(i.e. MSOA), while r s 1 describes the subsequent<br />

cooperative formation of ordered aggregates by a HG<br />

mechanism. Essentially, the treatment is a generalization<br />

of the Zimm-Bragg and the Oosawa theories (Kh A K) without<br />

necessarily specifying a detailed molecular model and<br />

a critical nucleus. With respect to Oosawa’s treatment that<br />

focused on the critical concentration C* (cf. Figure 4c),<br />

Figure 15. (a) Deoxyguanosine, its oligomers, and folic acid.<br />

Their assembly in tetrameric disks. [36] (b) Variation of the number<br />

of stacked tetrameric disks with folate concentration in (1)<br />

pure H2O and (2) 1 m NaCl at 308 C. The vertical broken line<br />

indicates the I e H transition (replotted using data taken from<br />

ref. [26] ).<br />

the van der Schoot treatment emphasizes the fractions of<br />

aggregated material and helical bonds as a function of<br />

both temperature and concentration. Rigidity and excluded<br />

volume effects are not introduced and, therefore, liquid<br />

crystallinity does not direct the aggregation of the stacks.<br />

5.3.3 Supermolecules. Tubular Assemblies<br />

The formation of disk-like supermolecules from two or<br />

more complementary components was discussed by several<br />

authors. [2, 34–36, 43, 44] Disk-like supermolecules often<br />

show liquid-crystalline behavior even though the separate<br />

components do not. Moreover, the discotic supermolecules<br />

can form columnar stacks in the melt and in solution<br />

just as the molecular discotics discussed above do.<br />

The relative contributions of MSOA, HG and SLC<br />

mechanisms have not always been characterized adequately.<br />

Gottarelli et al. [35, 36] have investigated the most<br />

interesting assembly of the nucleotide deoxyguanosine,<br />

its oligomers and alkaline folates (Figure 15a). These<br />

compounds form hydrogen-bonded disk-like tetramers in<br />

solution and are able to assemble in columnar stacks of<br />

discrete length and DP. Small-angle neutron scattering<br />

techniques were used to determine the length of the<br />

aggregates in water and in salt solutions. The critical concentrations<br />

for the appearance of the nematic (cholesteric)<br />

and the hexagonal phases were determined by<br />

means of X-ray diffraction. Selected data for the deoxyguanosine<br />

dimer (d(GpG); Figure 15a) and the folate is


<strong>Supramolecular</strong> <strong>Polymerizations</strong> 523<br />

Table 2. Critical concentration and DP in the isotropic phase<br />

for folic acid (selected data taken from ref. [36] ).<br />

Sample Solvent C IN<br />

%<br />

C NH<br />

%<br />

LISO DPISO a) XISO b)<br />

d(GpG) H2O – – 70 15 2.3<br />

d(GpG) H2O +Na + 2.5 18 – – –<br />

d(GpG) H2O +K + 1.5 15 – – –<br />

folate H2O – 35 2.3 1 0.1<br />

folate H2O +Na + 27 35 2.1 9 0.7<br />

a) L/2.35 Š (4.70 for d(GpG)).<br />

b) L/30 Š.<br />

collected in Table 2. The deoxyguanosine derivatives<br />

generally show larger DPs in isotropic solutions and<br />

lower critical concentrations CIN and CNH than the folates.<br />

The presence of NA + and, particularly, K + ions enhances<br />

the stabilization of the aggregates. In the case of folates<br />

in pure H2O, no nematic phase and small DPs were<br />

observed. Due to the considerable diameter of the cylinders<br />

(D L 30 Š) and the thickness of each disk (L = 2.35<br />

Š), the DP in the isotropic phase is extremely small and<br />

the corresponding axial ratio X suggests that thick disks<br />

rather than columns are present. Figure 15b illustrates the<br />

evolution of DP with concentration covering the range<br />

from the isotropic to the hexagonal phase. The smooth<br />

dependence DP vs C does not evidence cooperative<br />

effects in the case of folates. The largest DP (L30), determined<br />

from a 60% (hexagonal) solution, reveals in fact<br />

columns of very small geometric anisotropy (X L 2.3).<br />

The formation of the mesophase may be promoted by the<br />

large excluded volume effect of disks even in the absence<br />

of soft interactions. In fact, simulation studies evidenced<br />

nematic and columnar phases for solutions of extremely<br />

thin disks (0 a L/D a 0.1). [101]<br />

Disk-like supermolecules based on dimers of ureidotriazines<br />

connected by a 4 H-bond scheme similar (but<br />

not identical) to that of the ureidopyrimidone polymers in<br />

Figure 7c were reported by Meijer and coworkers. [34]<br />

These disks stack in columns with loose positional order<br />

and low DP (Figure 16a). Percec and coworkers [44]<br />

reported tubular polymeric assemblies of disks composed<br />

by six tapered molecules of 12-ABG-15C5 complexed<br />

with triflate salt (Figure 16b). The columns assembled<br />

into a hexagonal mesophase revealed as by means of Xray<br />

diffraction from the undiluted system.<br />

Several of the systems described above present a central<br />

cavity into which a covalent polymer can be hosted<br />

or a flow of ions be achieved (cf. also next section). Of<br />

particular interest are nanotubes (Figure 16c) formed by<br />

stacking cyclic peptides connected by H-bonds along the<br />

columnar axis. [37–41] The chemical design of these flat<br />

ring-like peptides was discussed by De Santis and coworkers.<br />

[37] Tubes assembled in solution and the contact<br />

forces for the dimerization (K L 2.5610 3 m –1 ) are too<br />

Figure 16. (a) Monofunctional ureidotriazine disks capable of<br />

assembling into columns. [34] (b) Assembly of tapered 12-ABG-<br />

12C5 into disks, formation of a column of stacked disk, hexagonal<br />

columnar organization. [44] (c) Nanotubules formed by Hbonded<br />

cyclic peptides. [37] (i Am. Chem. Soc. 1994 and 1996.)<br />

small for a large DP unless cooperative effects occur. [38]<br />

Cyclic b-peptides were also considered. [39] The self<br />

assembly of the nanotubules into ion-selective membranes<br />

was discussed as well. [40]<br />

Unimers of most of the systems considered above were<br />

subsequently connected by flexible covalent spacers, producing<br />

main-chain or side-chain SPs that should be<br />

described more appropriately under class C SPs. The<br />

basic ability of the disks to form columnar assemblies<br />

was preserved, but the covalent segments produced<br />

alterations in the stacking details such as the occurrence<br />

of helicity. For instance, a slowly rising helicoidal stack<br />

was produced when the crown ether receptor in Figure<br />

16b was replaced by a flexible endo-receptor (nEO-<br />

PMA) connected as a side-chain to a poly(methyl acrylate)<br />

chain. [42] When the ureidotriazine disks in Figure<br />

16a were main-chain linked through flexible spacers,<br />

helical columns and large DPs were observed. [34]<br />

Applications. The columnar nematic or hexagonal<br />

packing of disk-like molecules and supermolecules could<br />

be exploited as a precursor step for the assembly of large,<br />

oriented structures modeling natural systems. Electronic<br />

mobility due to the p–p interactions along the columnar<br />

axis may be useful for electronic and photonic<br />

devices. [102] Central cavities could be exploited for the<br />

selective hosting of polymer molecules [44] or for ion-<br />

[40, 41]<br />

selective channels.<br />

5.4 Host/Guest Polymeric Assemblies<br />

Covalent polymers can enter a cavity of a columnar<br />

assembly or of single ring-like structures. The result is a<br />

composite host/guest polymeric assembly exhibiting a


524 A. Ciferri<br />

Figure 17. Shish-kebab composites. Schemes for (a) polyrotaxane,<br />

[61] (b) tobacco mosaic virus, [50] and (c) a-cyclodextrin +<br />

poly(ethylene oxide) (taken from ref. [104] ).<br />

shish-kebab-type architecture. The driving force for the<br />

formation of these assemblies is a complex combination<br />

of molecular recognition and supramolecular polymerization.<br />

In fact, the host polymers often promote the supramolecular<br />

polymerization of the guest, or an alteration of<br />

its assembly mode. Three examples are illustrated in Figure<br />

17.<br />

The primary interaction assisting the threading of a<br />

polymer into a single macrocycle cavity, as in the case of<br />

pseudopolyrotaxanes (Figure 17a), is attributed to the<br />

occurrence of appropriately spaced p-rich hydroquinone<br />

rings on the polymer and p-acceptor groups within the tetracationic<br />

cyclophane. [59–61] It has been shown that the<br />

electron donor/acceptor interaction can be monitored by<br />

electrochemically or photochemically induced reduction/<br />

oxidation reactions. Relative motion of the two components<br />

can thus be induced, simulating a molecular microengine.<br />

[103]<br />

The situation of TMV, illustrated in Figure 17b, is<br />

more complex. Here the guest is an RNA molecule and<br />

the host is a hollow columnar assembly composed of<br />

2130 identical tapered protein molecules. The host/guest<br />

systems can be disassembled and reassembled in vitro by<br />

pH changes. However, the host can be reassembled even<br />

without RNA. A very interesting effect is manifested in<br />

the structure of the host when RNA is present. [50] In the<br />

absence of RNA the host is a stack of disks of various<br />

DPs, each disk comprising 17 protein units. However,<br />

formation of a spiral occurs when RNA occupies the cavity.<br />

The proteins of the host then follow a helical pattern<br />

with 16.3 units per turn, and the assembly assumes definite<br />

dimensions (L = 3000, d = 180 Š, X = 16.6) and a<br />

DP of 2310.<br />

The RNA-induced helix formation in an otherwise<br />

stacked systems of disks is reminiscent of the similar<br />

effect described in the preceding section (Figure 16a,b).<br />

It thus appears that supramolecular interactions between<br />

sites on RNA and protein induce spiral formation similar<br />

to that of disks connected to a covalent polymer as side<br />

chains. The complex role of RNA for the whole structure<br />

is evident. RNA acts like a crank-shaft that drives the<br />

proteins bound to it into a helical pattern and simultaneously<br />

provides the information about the proper length<br />

and DP of the host. The assembly mechanism of the overall<br />

TMV structure can thus be described in terms of a<br />

supramolecular polymerization of the columnar assembly,<br />

coupled to the formation of monofunctional sidechain<br />

bonds between RNA and proteins. [77]<br />

Figure 17c illustrates the assembly of a-cyclodextrin<br />

rings over poly(ethylene oxide). This system belongs to<br />

the class of inclusion compounds, or clathrates, that have<br />

aroused considerable interest for separation processes and<br />

for the unique properties of single chains confined in narrow<br />

(d L 6 Š) channels. [104] The stability of the crystalline<br />

adduct is likely to be assisted by pairwise host/guest<br />

interactions, the strength of which is increased within the<br />

small cavity. [56] However, the wide variety of systems<br />

capable of forming inclusion compounds invites to consider<br />

other less specific factors affecting the supramolecular<br />

polymerization of a-cyclodextrin rings threaded<br />

along the polymer chain. These factors might be: (i) relatively<br />

strong contact forces between the surfaces of the<br />

host and (ii) a steric-type effect not so far theoretically<br />

described. In support of (i) one may note that soluble stoichiometric<br />

complexes of the host are known to occur<br />

(e.g., head-to-head dimers of cyclodextrin unable to<br />

assemble into long channels in the crystalline structure).<br />

Concerning (ii) it is plausible that the guest stretches out<br />

(loss of conformation entropy) while simultaneously<br />

assembling individual host molecules. A suppression of<br />

undulation modes of the polymer due to the presence of<br />

rings may lead to an entropically induced effective attraction<br />

between the threaded rings (a similar Casimir-type<br />

effect leads to an attraction between undulating, flexible<br />

membranes). Single host/guest channels could form even<br />

in isotropic solution of more rigid polymers if there is a<br />

favorable balance between the contact energy of cyclodextrin<br />

rings and the chain-conformational entropy. In<br />

concentrated solutions, the resulting rod-like structure<br />

could be favored by the occurrence of a nematic phase.<br />

Applications. The possibility of generating relative<br />

motion of the assembled surfaces by electrochemical or<br />

photochemical stimuli could be exploited in a variety of


<strong>Supramolecular</strong> <strong>Polymerizations</strong> 525<br />

Figure 18. S-layers with hexagonal and square lattice symmetry<br />

derived from TEM. (a) Thermoanaerobacter thermohydrosulfuricus,<br />

and (b) Desulfotomaculum nigrificans (taken from<br />

ref. [46] ).<br />

nano/molecular scale engines. [103] The encapsulation of<br />

polymer molecules within cavities formed by self-assembling<br />

unimers provides systems of interest for separation<br />

processes, for recognizing and storing sequential information,<br />

[50] for orienting and screening single polymer molecules<br />

from similar neighbor interaction. [104]<br />

5.5 Planar Assemblies<br />

Figure 3c illustrates an equatorial distribution of four<br />

binding sites suitable for the formation of planar assemblies.<br />

As discussed in Section 4.2, these assemblies are<br />

expected to grow to large sizes by an intra-assembling<br />

cooperative mechanism akin to crystallization. At variance<br />

with the growth-coupled-to-orientation of linear<br />

systems, the growth of a planar assembly does not require<br />

the simultaneous formation and orientation of other growing<br />

units. Single free-standing, monomolecular layers are<br />

possible. An excellent verification of these expectations<br />

is provided by self-assembling S-layers forming the protective<br />

layer of the external surfaces of bacterial cells,<br />

and enabling the maintenance of a closed lattice during<br />

cell growth and division. [46] The identical constituent proteins<br />

have quasi-spherical form and exhibit an equatorial<br />

distribution of donor/acceptor groups capable of H-bonding<br />

to adjacent unimers. The proteins also posses a southpole<br />

site capable of electrostatic anchoring to the cell surface.<br />

S-layers can be disassembled and reassembled in<br />

vitro, allowing the preparation of purely H-bonded monolayers<br />

standing over an inert surface. The assembly q<br />

disassembly process has been described as a crystalliza-<br />

tion, [46] producing highly organized morphologies such as<br />

those shown in Figure 18. Depending upon the lattice<br />

type, the center-to-center distance of the morphological<br />

units varies from 3 to 30 nm, the thickness of monomolecular<br />

lattices vary from 5 to 25 nm, and the pore size is<br />

between 2 and 8 nm.<br />

Applications. The controllable confinement in definite<br />

areas of nanometric dimensions, coupled to the easiness<br />

of extraction and re-assembly, has allowed applications<br />

in areas of nanotechnology, such as bioanalytical sensors,<br />

templates for superlattices with prescribed symmetry,<br />

electronic and optical devices, matrices for the immobilization<br />

of functional molecules, and biocompatible surfaces.<br />

[46] S-layers recrystallized over solid supports have<br />

also been successfully patterned by using UV radiation<br />

and microlithographic masks. [46]<br />

5.6 Composite and 3D Assemblies<br />

Hexagonal cylindrical and lamellar phases (cf. transmission<br />

electron microscopy (TEM) photographs in Figure<br />

18) are often seen [48] in diblock and multiblock copolymers,<br />

ternary systems, copolymers formed from one unit<br />

that can be crystallized, rod-coil copolymers, [49] and some<br />

biological fibers. [105] For amorphous diblock copolymers<br />

in the cylindrical mode, one block is hexagonally packed<br />

within a matrix of the other block. The lamellar mode is<br />

instead based on alternating layers of A and B. The<br />

lamellar mode is the prevalent feature observed with rodcoil<br />

copolymers. [49, 106] In the case of a helical comb-like<br />

polymer (poly(b-l-aspartate) with paraffinic side chains),<br />

a layered distribution of helices correlated by interdigitation<br />

of the side chains was observed. [107] In the case of<br />

keratin, the fiber cross-section reveals L1 lm long microfibrils<br />

parallel to the fiber axis and hexagonally imbedded<br />

in a disordered S-rich matrix. Each microfibril is composed<br />

of eight protofibrils that are left-hand cables of two<br />

strands, each including two right-hand a-helices. [105]<br />

In the case of amorphous block copolymers, the (selfconsistent)<br />

mean-field theory [87] (cf. Section 4.2 and Figure<br />

6) describes the occurrence of various phases in terms<br />

of parameters pertinent to single copolymer molecules<br />

(compatibility, relative length and flexibility of the two<br />

blocks). This theory has been an eminently successful<br />

one and experimental results for the undiluted melt offer<br />

good support to it. [83, 108–110] Even in the case of block<br />

copolymer solutions, the predicted [88] sequence of phases<br />

upon increasing the concentration (e.g., isotropic e<br />

micellar e cubic e hexagonal e lamellar) revealed simi-<br />

[97, 98]<br />

larities with experimental data.<br />

Within the context of supramolecular polymerization it<br />

is however useful to explore alternative descriptions of<br />

the above structures in terms of a simpler, less sophisticated<br />

approach based on the concept of self-assembling<br />

of specifically designed building blocks. A similar con-


526 A. Ciferri<br />

Figure 19. (a) Micelles of coil-coil diblock copolymers in a<br />

selective solvent undergoing supramolecular polymerization.<br />

Hexagonally packed morphology as determined by means of<br />

TEM. (b) Bilayers of rod-coil diblock copolymers in a solvent<br />

selective for the coil block, growth directions within the plane of<br />

the layer and perpendicular to it are shown. Layered morphology<br />

[48, 112]<br />

as determined by means of TEM.<br />

cept has been used to describe solid arrays displaying<br />

complex and ordered structurizations such as interpenetrating<br />

nets. [111] The present author had suggested [1, 77] that<br />

the three-dimensional solid state morphologies described<br />

above, originating from molecular recognition of similar<br />

blocks, ought to be described in terms of supramolecular<br />

polymerization. The approach requires the identification<br />

of unimers with proper functionality and of their one- or<br />

multidimensional growth mechanism. [112] Relevant to this<br />

end are the considerations set forth in Section 4.2 regarding<br />

the separation of the modes of longitudinal and lateral<br />

growth.<br />

In particular, the formation of the hexagonal phase for<br />

copolymers that cannot be crystallized could be described<br />

as an essentially one-dimensional growth of micellar<br />

unimers according to the mechanism of growth-coupledto-orientation.<br />

By analogy with the processes illustrated<br />

in Figure 13, leading to the hierarchical sequences 5 and<br />

6, it has been suggested that the growth process schematized<br />

in Figure 19a describes the formation of the hexagonal<br />

phase of coil-coil block copolymers. [112] Here it is suggested<br />

that, as for conventional surfactants, a block copolymer<br />

micelle can assume an elongated form playing the<br />

role of a bifunctional unimer. Upon increasing the concentration,<br />

the latter undergoes linear growth simultaneously<br />

with the development of nematic orientation.<br />

This is followed by the hexagonal columnar phase (the<br />

intermediate nematic phase may not appear for suitable<br />

combinations of contact forces and persistence length) at<br />

even higher concentration. The hexagonal phase should<br />

be viewed as an embryo of the final morphology in the<br />

condensed state. The coiled segments may dangle over<br />

the lateral surface in a disordered fashion rather than<br />

interdigitate regularly. Note that while the approach discussed<br />

above does not have the predictive features of the<br />

mean-field theory, it does predict a role of the persistence<br />

length for the primary length scale, which is not predicted<br />

by the latter theory.<br />

The expectation that linear growth controls the formation<br />

of hexagonal columnar mesophases is not limited to<br />

coil-coil copolymers, or to systems with long molecular<br />

axes normal to the growth direction. Bifunctional unimers<br />

unable to grow along the lateral dimensions should in<br />

general be candidates for linear growth. In the case of<br />

keratin fibers, considering that the length of the microfibrils<br />

by far exceeds the length of constituent chains and<br />

falls in the range of the persistent length reported for<br />

similar systems, [1] it is plausible that the assembly of the<br />

fibril is also directed by the growth-coupled-to-orientation<br />

mechanism. The following hierarchical assembly<br />

sequence has therefore been suggested: extrusion of the<br />

low sulfur protein into extracellular fluids e head-to-tail<br />

assembly of shorter unimers into individual microfibrils<br />

to a length related to the persistence length with simultaneous<br />

orientation in the mesophase e stabilization of the<br />

microfibril by internal 1S1S1 bonding and the two non<br />

a-helical terminals e crosslinking of the sulfur-rich component<br />

in the narrow interfibrillar space. [112]<br />

The formation of lamellar structures could also be<br />

described qualitatively in terms of the self-assembly of<br />

specifically designed building blocks. Considering the<br />

case of the single lamella schematized in Figure 19b,<br />

growth akin to crystallization can occur along two perpendicular<br />

in-plane directions: functionality is larger than<br />

two. This mode of growth is at striking variance with the<br />

case of the uni-dimensional growth of cylindrical, bifunctional<br />

unimers considered above. It remains, however, to<br />

be considered how the ordered polymerization along the<br />

direction perpendicular to the lamellar plane is achieved.<br />

It is possible that, as disoriented lamellae grow, a critical<br />

axial ratio is attained at which purely hard interactions<br />

stabilize a nematic phase of the discotic type. This type<br />

of order has been experimentally evidenced during the<br />

graphitization of organic materials when large and growing<br />

planar rings are formed. [113] The coiling segments protruding<br />

from the lamellae may dangle over the surface, as<br />

in the case of cylindrical assemblies.<br />

An alternative assembling mode along the direction<br />

perpendicular to the lamellar plane could be based on<br />

attractive forces, or interdigitation, among the coiled segments.<br />

This interaction does not appear relevant to the<br />

class of copolymers considered above, but was evidenced<br />

for the comb-like poly(b-l-aspartate) with paraffinic side<br />

chains. This system showed nematic order based on clusters<br />

of helices correlated by interdigitation of partly molten<br />

side chains. [106] The polymer formed a complete 3D<br />

structure based on a layered organization of helices with<br />

side chains crystallized in a separate hexagonal lattice.<br />

Applications. Block copolymers are known to represent<br />

an important class of materials allowing a desirable blend


<strong>Supramolecular</strong> <strong>Polymerizations</strong> 527<br />

of properties of different polymers, while preventing the<br />

undesirable (incompatibility-driven) macrophase separation<br />

of unconnected components. [47–49, 106] Pre-assembly<br />

followed by the growth of specifically designed unimers<br />

could represent a novel strategy toward the fabrication of<br />

composite structures with a prescribed distribution of<br />

components.<br />

6 Topics for Further Investigation<br />

Several aspects of fundamental and applied character<br />

appear to need more extensive investigation. Here we<br />

restrict attention to the polymer-like properties of SPs<br />

and the basic polymerization mechanisms.<br />

A main problem is the assessment of DP. A characteristic<br />

feature of these systems is that DP is a function of<br />

concentration. However, the determination of DP at a<br />

given concentration becomes complicated when using<br />

conventional molecular-weight determinations, requiring<br />

extrapolations to infinite dilution. Measurements under<br />

theta conditions should be preferred. The assessment of<br />

DP by means of SEC is also questionable for kinetically<br />

unstable SPs that display weak contact forces. In the case<br />

of actin, a successful determination of the DP distribution<br />

was reported using electron microscopy. [28] A reliable<br />

determination of the complete DP/concentration dependence<br />

may require the evaluation of binding constants and<br />

the application of theoretical expressions valid for a specific<br />

growth mechanism. For instance, Equation (1) can<br />

be used in the low concentration regime when MSOA is<br />

expected to prevail. The analysis of data at higher concentration<br />

could allow the detection of cooperative contributions<br />

arising from the HG or SLC mechanisms.<br />

There are several other parameters that could be determined<br />

by an extension of conventional polymer physical<br />

chemistry. One is the characterization of the rigidity of<br />

the assembly. The usual determination of the persistence<br />

length from solution studies may not be a viable one in<br />

the present context. An alternative approach, suitable for<br />

large values of q, is based on flexural rigidity data<br />

extracted from the thermally driven fluctuation of shape<br />

or end-to-end distance. [114, 115] Calculation approaches are<br />

also possible. [116] Rheological characterization of processing<br />

parameters under shear and elongational flow should<br />

aim to the coupling of polymer viscoelasticity and bond<br />

lability. Analogies should be considered with the behavior<br />

of living polymers, [91] covalent networks exhibiting<br />

labile crosslinkages, including those formed in the<br />

oriented state. The assessment of the mechanical strength<br />

of supramolecular bonds, particularly for the linear Hbond<br />

materials discussed in Section 5.1.1, is another open<br />

topic of great importance, e.g., for the performance of<br />

supramolecularly extended covalent polymers. An evaluation<br />

of the most suitable method [117] for calculating the<br />

elastic constants of SPs is needed. The characterization of<br />

the difference in the properties of reversible and irreversible<br />

polymers and copolymers could be better assessed by<br />

comparative studies on a given SP and on its analog<br />

obtained by transforming the supramolecular main-chain<br />

bonds into covalent ones.<br />

Analysis and more detailed investigations of the<br />

growth mechanisms are also needed. The sudden occurrence<br />

of growth when the nematic phase appears ought to<br />

be documented for other SPs, the rigidity of which needs<br />

to be independently assessed. The theoretical reasons preventing<br />

the observation of a nematic phase for block<br />

copolymers and some other amphiphiles need to be clarified<br />

and experimentally verified. Systems for which the<br />

contact energy, or equilibrium constant, can be systematically<br />

altered (e.g. by altering the number of pairwise<br />

interactions at given functionality) should be considered<br />

for a more stringent test of the relationship between K<br />

and DP. In this context, recent work has shown that<br />

bonds based on DNA base pairing produce well-behaved<br />

SPs. [118] An analysis of model systems in which only the<br />

site distribution is altered could allow an assessment of<br />

the parameters controlling linear versus helical growth.<br />

The detailed analysis of the steps contributing to the formation<br />

of host/guest composites (Section 5.4), and other<br />

hierarchical processes (Section 5.3), should help to elucidate<br />

the scaling up from nanometric to mesoscopic<br />

assemblies.<br />

Acknowledgement: The author expresses his appreciation to<br />

Prof. Paul van der Schoot for clarifying discussions and constructive<br />

criticism.<br />

Received: March 23, 2002<br />

Revised: May 13, 2002<br />

Accepted: May 13, 2002<br />

[1] A. Ciferri, in: <strong>Supramolecular</strong> Polymers, A. Ciferri, Ed.,<br />

Marcel Dekker, New York 2000, p. 1.<br />

[2] J.-M. Lehn, in: <strong>Supramolecular</strong> Polymers, A. Ciferri, Ed.,<br />

Marcel Dekker, New York 2000, p. 615.<br />

[3] The Crystal as a <strong>Supramolecular</strong> Entity, G. R. Desiraju,<br />

Ed., Wiley, New York 1996.<br />

[4] L. Brunsveld, B. J. B. Folmer, E. W. Meijer, R. P. Sijbesma,<br />

Chem. Rev. 2001, 101, 4071.<br />

[5] P. S. Corbin, S. C. Zimmermann, in: <strong>Supramolecular</strong> Polymers,<br />

A. Ciferri, Ed., Marcel Dekker, New York 2000, p.<br />

147.<br />

[6] J. Pranata, S. G. Wierschke, W. L. Jorgensen, J. Am. Chem.<br />

Soc. 1991, 113, 2810.<br />

[7] J. Y. Lee, P. C. Painter, M. M. Coleman, Macromolecules<br />

1988, 21, 954.<br />

[8] C.-M. Lee, A. C. Griffin, Macromol. Symp. 1997, 117,<br />

281.<br />

[9] C. He, A. M. Donald, A. C. Griffin, T. Waigh, A. H.<br />

Windle, J. Polym. Sci. B 1998, 36, 1617.


528 A. Ciferri<br />

[10] C. B. St. Pourcain, A. C. Griffin, Macromolecules 1995,<br />

28, 4116.<br />

[11] P. Bladon, A. C. Griffin, Macromolecules 1993, 26, 6004.<br />

[12] M. Lee, B. K. Cho, Y. S. Kang, W. C. Zin, Macromolecules<br />

1999, 32, 8531.<br />

[13] J.-M. Lehn, Makromol. Chem., Macromol. Symp. 1993, 69,<br />

1.<br />

[14] M. Kotera, J.-M. Lehn, J.-P. Vigneron, Tetrahedron 1995,<br />

51, 1953.<br />

[15] M. Kotera, J.-M. Lehn, J.-P. Vigneron, J. Chem. Soc.,<br />

Chem. Commun. 1994, 197.<br />

[16] S. Choi, X. Li, E. E. Simanek, R. Akaba, G. M. Whitesides,<br />

Chem. Mater. 1999, 11, 684.<br />

[17] N. Kimizuka, T. Kawasaki, K. Hirata, T. Kunitake, J. Am.<br />

Chem. Soc. 1995, 117, 6360.<br />

[18] F. Gemiti, V. Giancotti, A. Ripamonti, J. Chem. Soc. 1968,<br />

757, 763.<br />

[19] C. Dammer, P. Maldivi, P. Terech, J. M. Guenet, Langmuir<br />

1995, 11, 1500.<br />

[20] U. Michelsen, C. A. Hunter, Angew. Chem. Int. Ed. 2000,<br />

29, 764.<br />

[21] W. M. Gelbart, A. Ben-Shaul, D. Roux, in: Micelles, Membranes,<br />

Microemulsions and Monolayers, Springer-Verlag,<br />

New York 1994.<br />

[22] T. Odijk, J. Phys. 1987, 48, 125.<br />

[23] R. Hentschke, B. Fodi, in: <strong>Supramolecular</strong> Polymers, A.<br />

Ciferri, Ed., Marcel Dekker, New York 2000, p. 61.<br />

[24] G. Porte, Y. Poggi, J. Appell, G. Maret, J. Phys. Chem.<br />

1984, 88, 5713.<br />

[25] F. Oosawa, in: <strong>Supramolecular</strong> Polymers, A. Ciferri, Ed.,<br />

Marcel Dekker, New York 2000, p. 643.<br />

[26] F. Oosawa, S. Asakura, in: Thermodynamics of the Polymerization<br />

of Protein, Academic Press, London 1975, p. 25.<br />

[27] E. D. Korn, Physiol. Rev. 1982, 62, 672.<br />

[28] P. A. Janmei, J. Peetermans, K. S. Zanert, T. P. Stossel, T.<br />

Tanaka, J. Biol. Chem. 1986, 261, 8357.<br />

[29] F. Oosawa, Biophys. Chem. 1993, 47, 1010.<br />

[30] M. O. Steinmetz, D. Stoffler, A. Hoenger, A. Bremer, U.<br />

Aebi, J. Struct. Biol. 1997, 119, 295.<br />

[31] R. Hentschke, R. Edwards, P. J. B. Boden, R. J. Bushby,<br />

Macromol. Symp. 1994, 81, 361.<br />

[32] R. A. Palmans, J. A. J. M. Vekemans, E. E. Havinga, E. W.<br />

Meijer, Angew. Chem. Int. Ed. Engl. 1997, 36, 2648.<br />

[33] L. Brunsveld, H. Zhang, M. Glasbeek, J. A. J. M. Vekemans,<br />

E. W. Meijer, J. Am. Chem. Soc. 2000, 122, 6175.<br />

[34] J. H. K. K. Hirschberg, L. Brunsveld, A. Ramzi, J. A. J. M.<br />

Vekemans, R. P. Sijbesma, E. W. Meijer, Nature 2000,<br />

407, 167.<br />

[35] G. Gottarelli, G. P. Spada, A. Garbesi, in: Comprehensive<br />

<strong>Supramolecular</strong> Chemistry, J.-M. Lehn, Ed., Pergamon<br />

Press, Oxford 1996, Vol. 9, p. 483.<br />

[36] G. Gottarelli, G. P. Spada, P. Mariani, in: Crystallography<br />

of <strong>Supramolecular</strong> Compounds, G. Tsoucaris, J. L.<br />

Atwood, J. Lipkowski, Eds., Kluwer Academic Publishers<br />

1996, p. 307.<br />

[37] P. De Santis, S. Morosetti, R. Ricco, Macromolecules<br />

1974, 7,52.<br />

[38] D. H. Lee, M. R. Ghadiri, in: Comprehensive <strong>Supramolecular</strong><br />

Chemistry, J.-M. Lehn, Ed., Pergamon, New York<br />

1996, Chapter 12.<br />

[39] D. Seebach, J. L. Matthews, A. Meden, T. Wessels, C. Baerlocher,<br />

L. B. McCusker, Helv. Chim. Acta 1997, 80, 173.<br />

[40] H. S. Kim, J. D. Hartgerink, M. R. Ghadiri, J. Am. Chem.<br />

Soc. 1998, 120, 4417.<br />

[41] K. Motesharei, M. R. Ghadiri, J. Am. Chem. Soc. 1997,<br />

119, 11306.<br />

[42] V. Percec, J. Heck, G. Johansen, G. Tomazos, M. Kawagumi,<br />

J. Macromol. Sci.-Pure Appl. Chem. 1994, A11,<br />

1031.<br />

[43] V. Percec, C.-H. Ahn, G. Ungar, D. J. P. Yeardly, M. Möller,<br />

Nature 1998, 391, 161.<br />

[44] V. Percec, G. Johansson, G. Ungar, J.P. Zhou, J. Am.<br />

Chem. Soc. 1996,118, 9855.<br />

[45] C. Viney, Supramol. Sci. 1997, 4, 75.<br />

[46] U. B. Sleytr, M. Sàra, D. Pum, in: <strong>Supramolecular</strong> Polymers,<br />

A. Ciferri, Ed., Marcel Dekker, New York 2000, p.<br />

177.<br />

[47] A. Gabellini, M. Novi, A. Ciferri, C. Dell’Erba, Acta<br />

Polym. 1999, 50, 127.<br />

[48] V. Abetz, in: <strong>Supramolecular</strong> Polymers, A. Ciferri, Ed.,<br />

Marcel Dekker, New York 2000, p. 215.<br />

[49] K. Loos, S. Muæoz-Guerra, in: <strong>Supramolecular</strong> Polymers,<br />

A. Ciferri, Ed., Marcel Dekker, New York 2000, p. 263.<br />

[50] A. Klug, Angew. Chem., Int. Ed. Engl. 1983, 22, 565.<br />

[51] H. Kihara, T. Kato, T. Uryu, J. M. J. FrØchet, Chem. Mater.<br />

1996, 8, 961.<br />

[52] R. P. Sijbesma, F. H. Beijer, L. Brunsveld, B. J. B. Folmer,<br />

J. H. K. K. Hirschberg, R. F. M. Lange, J. K. L. Lowe, E.<br />

W. Meijer, Science 1997, 278, 1601.<br />

[53] M. V. Meftahi, J. M. J. FrØchet, Polymer 1988, 29, 477.<br />

[54] D. J. Massa, K. A. Shriner, S. R. Turner, B. I. Voit, Macromolecules<br />

1995, 28, 3214.<br />

[55] R. F. M. Lange, E. W. Meijer, Macromolecules 1995, 28,<br />

782.<br />

[56] H.-J. Schneider, Angew. Chem. Int. Ed. Engl. 1991, 30,<br />

1417.<br />

[57] C. M. Paleos, J. Michas, Liq. Cryst. 1992, 11, 773.<br />

[58] T. Kato, J. M. J. FrØchet, Macromol. Symp. 1995, 98, 311.<br />

[59] M. C. T. Fyfe, J. F. Stoddart, Acc. Chem. Res. 1997, 30,<br />

393.<br />

[60] F. M. Raymo, J. F. Stoddart, in: <strong>Supramolecular</strong> Polymers,<br />

A. Ciferri, Ed., Marcel Dekker, New York 2000, p. 323.<br />

[61] D. Philp, J. F. Stoddart, Synlett 1991, 7, 445.<br />

[62] D. A. Tomalia, I. Majoros, in: <strong>Supramolecular</strong> Polymers,<br />

A. Ciferri, Ed., Marcel Dekker, New York 2000, p. 359.<br />

[63] F. O’Brien, Trends Polym. Sci. 1994, 2, 183.<br />

[64] G. Liu, L. Qiao, A. Guo, Macromolecules 1996, 29, 5508.<br />

[65] W. Stocker, B. Karakaya, L. B. Schurmann, P. J. Rabe,<br />

A.-D. Schlüter, J. Am. Chem. Soc. 1998, 120, 7691.<br />

[66] G. Johansson, V. Percec, G. Ungar, J. P. Zhou, Macromolecules<br />

1996, 29,646.<br />

[67] V. Percec, D. Schlüter, Macromolecules 1997, 30, 5783.<br />

[68] R. Yin, Y. Zhu, D. A. Tomalia, J. Am. Chem. Soc. 1998,<br />

120, 2678.<br />

[69] L. Yan, W. T. S. Huck, G. M. Whitesides, in: <strong>Supramolecular</strong><br />

Polymers, A. Ciferri, Ed., Marcel Dekker, New York<br />

2000, p. 435.<br />

[70] X. Arys, A. M. Jonas, A. Laschewsky, R. Legras, in:<br />

<strong>Supramolecular</strong> Polymers, A. Ciferri, Ed., Marcel Dekker,<br />

New York 2000, p. 505.<br />

[71] J. Rühe, W. Knoll, in: <strong>Supramolecular</strong> Polymers, A.<br />

Ciferri, Ed., Marcel Dekker, New York 2000, p. 565.<br />

[72] P. Hobza, R. Zahradnik, Intermolecular Complexes, Elsevier,<br />

Amsterdam 1988.<br />

[73] G. Odian, Principles of Polymerization, McGraw-Hill,<br />

New York 1991.<br />

[74] H. Sund, K. Markau, Int. J. Polym. Mater. 1976, 4, 251.<br />

[75] R. Bruce Martin, Chem. Rev. 1996, 96, 3043.<br />

[76] J. van Gestel, P. van der Schoot, M. A. J. Michels, J. Phys.<br />

Chem. B. 2001, 105, 10691.<br />

[77] A. Ciferri, Liq. Cryst. 1999, 26, 489.<br />

[78] R. Hentschke, Liq. Cryst. 1991, 10, 691.


<strong>Supramolecular</strong> <strong>Polymerizations</strong> 529<br />

[79] P. van der Schoot, J. Phys. II 1995, 5, 243.<br />

[80] J. N. Israelachvili, in: Intermolecular and Surface Forces,<br />

Academic Press, London 1992, p. 341.<br />

[81] P. van der Schoot, J. Phys. Chem. 1992, 96, 6083.<br />

[82] A. Ciferri, Macromol. Chem. Phys. 1994, 195, 457.<br />

[83] T. Hashimoto, N. Sakamoto, T. Koga, Phys. Rev. 1996,<br />

E54, 5832.<br />

[84] E. Helfand, J. Chem. Phys. 1975, 62, 999.<br />

[85] L. Leibler, Macromolecules 1980, 13, 1602.<br />

[86] N. Semenov, Sov. Phys. JETP 1985, 61, 733.<br />

[87] M. W. Matsen, F. S. Bates, Macromolecules 1996, 29,<br />

1091.<br />

[88] J. Noolandi, A.-C. Shi, P. Linse, Macromolecules 1996,<br />

29, 5907.<br />

[89] R. K. Castellano, R. Clark, S. L. Craig, C. Nuckolls, J.<br />

Rebek, Jr., Proc. Natl. Acad. Sci. USA 2000, 97, 12418.<br />

[90] H. Hilger, R. Stadler, Macromolecules 1992, 25, 6670.<br />

[91] M. E. Cates, S. J. Candau, J. Phys. B: Condens. Matter<br />

1990, 2, 6869.<br />

[92] A. Ciferri, R. Kotex, J. Preston, work in progress.<br />

[93] A. Ciferri, Trends Polym. Sci. 1997, 5, 142.<br />

[94] T. Odijk, Curr. Opin. Colloid Interface Sci. 1996, 1, 337.<br />

[95] P. van der Schoot, M. E. Cates, Europhys. Lett. 1994, 25,<br />

515.<br />

[96] T. Shikata, D. S. Pearson, Langmuir 1994, 10, 4027.<br />

[97] K. Zhang, A. Khan, Macromolecules 1995, 28, 3807.<br />

[98] M. Svensson, P. Alexandridis, P. Linse, Macromolecules<br />

1999, 32, 637.<br />

[99] P. Linse, J. Phys. Chem. 1993, 97, 13896.<br />

[100] P. van der Schoot, M. A. J. Michels, L. Brunsveld, R. P.<br />

Sijbesma, A. Ramzi, Langmuir 2000, 16, 10076.<br />

[101] M. A. Bates, D. Frenkel, Phys. Rev. E 1998, 57, 4824.<br />

[102] M. Van de Craats, J. M. Warman, A. Fechtenkötter, J. D.<br />

Brand, M. A. Harbison, K. Müllen, Adv. Mater. 1999, 11,<br />

1469.<br />

[103] V. Balzani, A. Credi, F. M. Raymo, J. F. Stoddart, Angew.<br />

Chem. Int. Ed. 2000, 39, 3348.<br />

[104] L. Huang, A. E. Tonelli, J. Macromol. Sci., Rev. Macromol.<br />

Chem. Phys. 1998, C38, 781.<br />

[105] M. K. Hartzer, Y.-Y. S. Pang, R. M. Robson, J. Mol. Biol.<br />

1985, 365, 376.<br />

[106] J. T. Chen, E. L. Thomas, C. K. Ober, S. S. Hwang,<br />

Macromolecules 1995, 28, 1688.<br />

[107] F. Lopez-Carrasquero, S. Monserrat, A. Martinez de<br />

Harduya, S. Muæoz-Guerra, Macromolecules 1995, 28,<br />

5535.<br />

[108] F. S. Bates, M. F. Schult, A. K. Khandpur, S. Förster, J.<br />

H. Rosedale, K. Almdal, K. Mortensen, Faraday Discuss.<br />

1994, 98, 7.<br />

[109] S. Koizumi, H. Hasegawa, T. Hashimoto, Macromolecules<br />

1994, 27, 4371.<br />

[110] K. L. Winey, E. L. Thomas, L. J. Fetters, Macromolecules<br />

1992, 25, 422.<br />

[111] S. R. Batten, R. Robson, Angew. Chem. Int. Ed. 1998, 37,<br />

1460.<br />

[112] A. Ciferri, Recent Res. Dev. Macromol. Res., in preparation.<br />

[113] L. S. Singer, in: Ultra High Modulus Polymers, A.<br />

Ciferri, I. M. Ward, Eds., Applied Science Publishers,<br />

London 1979, p. 251.<br />

[114] L. D. Landau, E. M. Lifshitz, Statistical Physics, Pergamon<br />

Press, Tarrytown, N.Y. 1980.<br />

[115] F. Gittes, B. Mickey, J. Nettleton, J. Howard, J. Cell.<br />

Biol. 1993, 120, 923.<br />

[116] F. LauprÞtre, C. No l, Liquid Crystallinity in Polymers,<br />

A. Ciferri, Ed., VCH, Weinheim 1991, p. 3.<br />

[117] G. Capaccio, I. M. Ward, Polym. Eng. Sci. 1975, 15, 219.<br />

[118] E. A. Fogleman, V. R. Kempf, S. L. Craig, Polym. Prepr.<br />

(Am. Chem. Soc., Div. Polym. Chem.) 2002, 43, 568.

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!