The future of supramolecular copolymers unveiled by reflecting on

7 days ago - Supramolecular copolymers are an emerging class of mate-rials, and in the last years their potential has been demonstrated on a broad sca...
0 downloads 0 Views 3MB Size
Subscriber access provided by READING UNIV

Perspective

The future of supramolecular copolymers unveiled by reflecting on covalent copolymerization Beatrice Adelizzi, Nathan Van Zee, Lafayette de Windt, Anja R.A. Palmans, and Egbert Willem Meijer J. Am. Chem. Soc., Just Accepted Manuscript • DOI: 10.1021/jacs.9b01089 • Publication Date (Web): 19 Mar 2019 Downloaded from http://pubs.acs.org on March 20, 2019

Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.

is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.

Page 1 of 21 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Journal of the American Chemical Society

The future of supramolecular copolymers unveiled by reflecting on covalent copolymerization Beatrice Adelizzi,1,2 Nathan J. Van Zee,1,2,3 Lafayette N. J. de Windt,1,2 Anja R. A. Palmans,1,2 * E. W. Meijer1,2 * 1 Laboratory of Macromolecular and Organic Chemistry, 2 Institute for Complex Molecular Systems, Department of Chemical Engineering and Chemistry, Eindhoven University of Technology, P.O. Box 513, 5600 MB Eindhoven, The Netherlands, 3 Chimie Moléculaire, Macromoléculaire, et Matériaux, École Supérieure de Physique et de Chimie Industrielles de la Ville de Paris (ESPCI)–CNRS, UMR-7167, Paris Sciences et Lettres (PSL) Research University, 10 Rue Vauquelin, 75005 Paris, France

ABSTRACT: Supramolecular copolymers are an emerging class of materials, and in the last years their potential has been demonstrated on a broad scale. Implementing noncovalent polymers with multiple components can bring together useful features such as dynamicity and new functionalities. However, mastering and tuning the microstructure of these systems is still an open challenge. In this perspective we aim to trace the general principles of supramolecular copolymerization by analyzing them through the lens of the well-established field of covalent copolymerization. Our goal is to delineate guidelines to classify and analyze supramolecular copolymers in order to create a fruitful platform to design and investigate new multicomponent systems.

connections helped to highlight the distinctions between covalent and supramolecular polymerization processes, rendering underlying research needs into tangible goals.

 INTRODUCTION

In this context, one of the directions that we identified was the intriguing challenge to make well-defined supramolecular copolymers. At that time, in contrast to the multitude of methods available to tailor covalent copolymers, there were few synthetic strategies to control the microstructure of supramolecular copolymers. The challenge is the inherent lability of non-covalent bonds, which are typically kinetically and thermodynamically less stable than covalent bonds. However, in the past ten years, considerable inroads have been made to reconcile the dynamicity of supramolecular copolymers with the demand for microstructural control. Many studies and comprehensive reviews2-6 have since been published focusing on nucleationelongation processes, kinetic control, competition, pathway complexity, and supramolecular monomer design.

While one-dimensional supramolecular polymers have long been employed by Nature as a powerful structural motif, scientists have only recently begun to exploit their dynamicity, directionality, and functionality. This confluence of properties makes them promising for applications in pharmaceuticals, nanoelectronics, and catalysis, but many fundamental principles that dictate the structure of supramolecular polymers remain unresolved. In 2009, we proposed to draw inspiration for future research directions by carefully considering the maturation of its older sister, covalent polymerization.1 Following Carothers’s classic classification, we organized supramolecular polymers based on the process through which each aggregate is formed. Connections between the two fields became strikingly clear. Isodesmic polymerization can be considered a non-covalent analogue of step growth polymerization, characterized by high polydispersity and a degree of polymerization that depends on the association constant of the supramolecular units. Ring–chain supramolecular polymerization is similar to ring opening polymerization and depends on the equilibrium between linear and cyclic aggregates. Finally, cooperative supramolecular polymerization can be correlated to chain growth polymerization. This process is typically under thermodynamic control; although kinetically trapped copolymers are reported, over time such species generally convert into the thermodynamically most stable state. These

Given the field’s nascence, we feel it is timely to critically reflect on recent progress through the lens of covalent chain copolymerization to illuminate challenges and opportunities for the future. Considering the ubiquity of well-defined covalent copolymers, we believe that controlling the structure of one-dimensional supramolecular copolymers is paramount to realize next-generation functional materials. Our objective is to generate insight into urgent research directions by creating a strong link between covalent and non-covalent copolymerization worlds. The first supramolecular copolymers, such as the one reported by Lehn and co-workers,7 are based on an isodesmic mechanism and non-covalent interactions among complementary units, but in this work we mainly focus on cooperative chain growth processes. In isodesmic (step) copolymerization, the sequence of repeating units is either dictated by differing complementarity or randomized in the case of equal reactivity. For cooperative (chain) copolymerization, the challenges are similar to those of their covalent counterparts. As a point of reference, we chose to structure this work based on the organization employed in the chain copolymerization chapter of Odian’s classic textbook Principles of Polymerization.8 We focus on the classification and synthesis of state-of-the-art structures, discussing the parallel roles of direct characterization and theoretical models that are fundamental for the development of the field.

ACS Paragon Plus Environment

Journal of the American Chemical Society 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

 SUPRAMOLECULAR CHAIN GROWTH COPOLYMERIZATION Covalent and non-covalent copolymerization: Similarities and differences. The mechanisms of covalent chain growth copolymerization and cooperative supramolecular (chain growth) copolymerization carry several similitudes. At the outset of both processes, a certain activation is required to initiate polymer growth. For covalent polymers, a reactive initiator (e.g., a radical, cation, or anion) reacts with a monomer to generate an active propagating species that enchains additional monomers (Figure 1a).8 Similarly, supramolecular polymers require the generation of a nucleus (e.g., generally an energetically unfavorable small aggregate) to initiate the cooperative growth of a supramolecular copolymer (Figure 1b).1 Just as in covalent processes, the initiation and growth of supramolecular copolymers are sensitive to impurities,9,10 cosolvents,11-13 reaction time, and temperature. All these considerations dictate average molecular weight and microstructure.14,15 Just as in covalent copolymerization, changes to viscosity15-17 and optical properties17-20 of a supramolecular polymer solution or melt signal the progress of reaction. Moreover, several synthetic strategies used to control structure in both covalent and supramolecular chemistries share mechanistic similarities. For example, the use of chain transfer agents and chain stoppers provides control over the average length of covalent and supramolecular polymers, respectively.21-23 In addition, the recent introduction of supramolecular living polymerization24-30 has unlocked control over copolymer length and sequence analogous to living chain copolymerization. However, the nature of the bonds that hold supramolecular and covalent polymers together are sharply distinct. In contrast to the thermodynamic stability of covalent copolymers, supramolecular copolymers are dynamic because they are held together by comparatively weak non-covalent interactions (e.g., hydrogen bonding, π-π stacking, and van der Waal’s forces). The monomer sequence of a supramolecular copolymer is fluxional, as monomers can change their relative positions within a given polymer chain or integrate into other chains. Consequently, supramolecular polymers do not terminate in the way that covalent polymers do. These properties impart unique adaptability and modularity to the resulting materials, but they render the analysis of their structure a great challenge.

Figure 1: Growth mechanisms for chain polymerization. a) Covalent polymer activated by an initiator, which subsequently undergoes propagation or termination. b)

Page 2 of 21

Supramolecular polymer formed by a cooperative mechanism, which is activated by the unfavorable formation of a nucleus and elongated by the favorable incorporation of additional monomers. Another striking difference is that supramolecular polymerizations often show pathway complexity.31 Small differences in the conditions of aggregation (e.g., variation in temperature or solvent) can cause a given set of monomers to form different self-assembled structures.32-34 Covalent polymerizations are also sensitive to changes in reaction conditions, giving rise to fluctuations in, for example, polydispersity, tacticity, and monomer sequence. However, excluding vitrimers and other reversible polymers, these changes are permanently encoded into the final covalent polymers. In the case of pathway complexity, as a consequence of the inherent dynamicity of supramolecular polymers, the concentration of each type of aggregate is not constant and responds to changes in reaction conditions. The ability to tune aggregated structures at will highlights the extreme sensitivity and potential utility of these structures.35 The non-innocent effect of solvents in supramolecular polymerization. Before proceeding, it is important to consider solvent in defining copolymer compositions. In the case of covalent copolymerization, the choice of solvent can be an important tool for tuning relative polymerization rates, which ultimately determines the microstructure of a copolymer. The interaction between the solvent and the propagating polymer species is frequently supramolecular in nature. For example, solvent polarity dictates the strength of ion pairing in the anionic copolymerization of vinyl monomers, which modulates the reactivity of the propagating carbanion chain-end toward different monomers depending on their respective structures.8 The resulting polymers are then typically represented in terms of their polymerized monomer units. When dispersed in a solvent, the solvent interacts with the polymer to some degree depending on the temperature and the reactivities of the polymer and the solvent. However, these associated solvent molecules are generally not included in the polymer’s chemical structure because they are not covalently bonded to the polymer. This distinction does not exist in the case of supramolecular copolymerization because solvent molecules interact with supramolecular copolymers via the same kinds of non-covalent forces that hold the monomers together. For this reason, a supramolecular “homopolymerization” of a single monomer may not be a homopolymerization at all. Solvent molecules could serve an essential role in supporting the fiber’s supramolecular structure, and thus such a naming convention is potentially misleading because it deemphasizes the importance of solvent molecules as structural units. It is nevertheless difficult to unambiguously assign the role of solvent in most supramolecular systems, and so we acknowledge the necessity of simplifying supramolecular polymer structures by describing them primarily in terms of their principle monomer component(s). Solvent effects have been reviewed elsewhere,36,37 but here we would like to highlight recent efforts to exploit the action of solvents as a means to control supramolecular structure. The most common manifestation of this approach is to add a “good” solvent (i.e., a solvent that competitively associates with the supramolecular units) to denature

ACS Paragon Plus Environment

Page 3 of 21 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Journal of the American Chemical Society

supramolecular fibers, thereby controlling their average size. Aida and co-workers38 recently expanded this principle by balancing denaturation with the self-association of the denaturant to realize bisignate supramolecular polymerization as a function of temperature. Others have shown that the helicity of one-dimensional fibers—usually evaluated using circular dichroism spectroscopy—can be controlled through tuning the interplay of structure and reactivity between both the supramolecular polymer and the solvent. Our group and others39,40 have shown that the shape of the solvent can directly influence the helicity of supramolecular fibers, suggesting that molecular packing of solvent into the fiber’s corona gives rise to perturbations to the inner core of the fiber.

Alternating copolymers, represented as poly(A-alt-B), contain the monomers in a regular alternating pattern. A related microstructure is periodic, which has the monomers sequenced in a regular pattern; for example, poly(A2-per-B) indicates that two A monomers are enchained in a row followed by one B monomer. Block and blocky copolymers contain both long sequences of each monomers in the copolymer chain. Diblock copolymers, poly(A-b-B), are linear polymers with long sequences of each monomers, while multi-block (“blocky”) copolymers exhibit multiple short runs of each monomer in a given polymer. As presented in Figure 2c, we classify supramolecular structures using the same IUPAC formalism that is used for covalent polymers.

Water must also be acknowledged for its importance as a structural element in supramolecular polymers in bulk and in solution. In the bulk, Qi, Schalley, and co-workers41 demonstrated that water acts as an essential structural component that unlocks impressive adhesive properties of supramolecular polymers formed with crown etherfunctionalized benzene tricarboxamides. In aqueous media, Han, Stupp and co-workers42 establish a striking resemblance between the interaction of water and synthetic peptide nanofibers with the interaction between water and natural supramolecular materials such as proteins. Water in the interior of such fibers exhibit rapid diffusion, while water at the surface of the fiber is essentially immobilized, suggesting that these water molecules form hydrogen bonds with the fiber and serve an integral role in supporting the fiber’s structure. In organic solvents as well, several reports43–49 point to the fundamental relationship between supramolecular structure and the presence of water—even trace amounts in highly apolar solvents. Our group48 recently reported that the monomeric nature of water in alkanes gives rise to a potent thermodynamic driving force for it to bind to co-dissolved supramolecular polymers and effect profound changes to their structures. We believe that this growing body of research expresses both the complexity and import of solvent molecules as not merely spectators but as active participants in one-dimensional supramolecular polymerization.

In the next sections, we critically review the theoretical tools and synthetic strategies for constructing well-defined one-dimensional structures. We shall consider the range of known microstructures and the limits of structural characterization. We focus especially on synthetic strategies and characterization challenges.

Type of copolymers. Linear covalent copolymers are commonly categorized by the pattern of monomer enchainment. For example, for a polymer made up of monomers A and B, we can assign it as random, alternating, block, blocky, or periodic (Figure 2). Random copolymers do not have a regular pattern of monomer enchainment.

 SUPRAMOLECULAR COPOLYMER COMPOSITION As in covalent copolymerization, theoretical models are essential for understanding the composition of supramolecular copolymers. Theory is especially valuable here because the inherent dynamicity of supramolecular copolymers makes it difficult to experimentally determine microstructure. In the model developed by ten Eikelder and Markvoort,50 this behavior is accounted for by treating supramolecular copolymerizations as a set of reversible reactions between monomers and polymers. Similar to the terminal model of chain copolymerization, the equilibrium constant of a reaction is assumed to depend only on the identity of the last-enchained monomer; an added assumption is that monomers can only add to the end of a growing chain. For example, the co-assembly of A and B to form supramolecular copolymer P can thus be described by four reactions (Figure 3a) that provide the basis for deriving expressions for the concentration and composition of the copolymers.

ACS Paragon Plus Environment

Journal of the American Chemical Society 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 4 of 21

Figure 2. Covalent and supramolecular chain copolymers. General mechanism for a) covalent and b) supramolecular chain copolymerizations. c) Possible microstructures for each type of copolymerization. Polymers are ordered from no inter-reactivity (blend for covalent and self-sorted for supramolecular polymers) to high inter-reactivity (alternate copolymers). The concentration of copolymers is expressed in terms of the concentration of polymerized monomer units ([M]pol), which in turn can be expressed in terms of the total concentration of monomer ([M]tot) and the concentration of free monomer ([M]free, Figure 3b, top). A similar equation can be derived for the fraction of polymerized monomers pol (Figure 3b, bottom). To determine the concentration of free monomer as well as the concentration and composition of the copolymers, the elaborated forms of these mass balance equations are solved numerically with the input consisting of the known [M]tot and the experimentally determined equilibrium constants KA-A, KA-B, KB-A, and KB-B (Figure 3a). From this model and subsequent fitting, we can extract thermodynamic parameters that help us to understand supramolecular microstructure. As in covalent copolymerization, reactivity ratios are a useful tool for comparing microstructures when the monomers are less, equally, or more reactive in copolymerization compared to homopolymerization. Here, the reactivity ratio R is defined as the ratio of Gibbs free energy change of the homoaggregation (ΔGA−A + ΔGB−B) to the Gibbs free energy change of the hetero-aggregation (ΔGA−B + ΔGB−A, Figure 3c). For values of R greater than unity, monomers preferentially interact with their own type of monomer, while for values of R smaller than unity, the monomers tend to co-assemble.

This behavior is reflected in the shape of the copolymerization curve obtained for each class of R. As an example, we simulated the results for the temperatureinduced supramolecular copolymerization of A and B. We arbitrarily set the thermodynamic parameters such that the elongation temperatures for poly(A) and poly(B) are 60 and 70 °C, respectively. We then varied the value of R by keeping the values of ΔGA−A and ΔGB−B constant and modifying the values of ΔGA−B and ΔGB−A. In this way, we modulate the tendency of the monomers to interact with each other while holding their tendency to homopolymerize constant. The resulting copolymerization curves (shades of grey), as well as the fraction of polymerized A and B (red and blue, respectively), are presented in Figure 3d. As R approaches infinity, two elongation temperatures are clearly evident, consistent with self-sorted homopolymerization of A and B. As the value of R is reduced, only one elongation temperature is evident, suggesting some degree of copolymerization. This elongation temperature shifts toward higher temperatures as the value of R is decreased—that is, the copolymer forms at higher temperature as heteroaggregation becomes more favorable. The incorporation of A and B in poly(A-co-B) depends on R as well. A coinciding incorporation of the two monomers is visible for alternating systems (R < 1) whereas increasing the value of R results in a greater disparity between the red and blue curves.

ACS Paragon Plus Environment

Page 5 of 21 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Journal of the American Chemical Society

Figure 3. Supramolecular copolymerization model. a) Terminal model of the supramolecular copolymerization between two types of monomers. The model assumes that the equilibrium constant depends only on the identity of the terminal monomer unit (PnA or PnB). b) The mass balance defines the concentration of polymerized monomer [M]pol as the difference between the total concentration of monomer [M]tot and the free monomer [M]free. The fraction of polymerized monomer is defined as pol. c) The composition and microstructure of the copolymer is determined by the reactivity ratio R, which is defined by the ratio of Gibbs free energy change of the homo-aggregation (ΔGA−A + ΔGB−B) to the Gibbs free energy change of the hetero-aggregation (ΔGA−B + ΔGB−A. d) Copolymerization curves are reported as a function of R: from left to the right, descending order from R approaching infinite values for self-sorted systems to R smaller than 1. e) Equation defining the fraction of A in the polymer (FA). f) Plot of FA vs. temperature as a function of R. A useful extrapolation is the calculation of the fraction of polymerized A, FA, as a function of temperature (Figure 3e and f). For R > 1 (black curve), at high temperatures, the copolymers are rich in monomers of type A. As temperature decreases, monomers of type B enter into the copolymers, and the composition of the copolymers approaches the composition of the mixture of monomers. For R < 1 (lightest grey curve), where alternating structures can be formed, at high temperatures monomers B are already incorporated into the copolymers and the copolymers’ composition already resembles the composition of the mixture. Thus, as the value of R decreases, monomers enter into the copolymers more by interacting with the other type of monomer and less by interacting with their own type of monomers. An important issue to keep in mind, as with all models, is that care must be taken to evaluate the validity of the assumptions for the system at hand. This model strictly considers the interactions between two monomers in a fully reversible systems—complications due to solvent interactions, pathway complexity, and kinetic traps are excluded. In cases where such factors strongly influence selfassembly, one can either modify the model to account for new interactions or use the base model without fitting to generate simulations for qualitative comparison.

 MICROSTRUCTURE OF COPOLYMERS Sequence-length distribution. For the sake of simplicity, the average copolymer microstructure is usually discussed. However, as indicated by the terminal model for covalent copolymerization and the thermodynamic model presented above for supramolecular copolymerization, the composition of a given copolymerization—and thus its precise microstructure—changes as a function of conversion. For covalent copolymers, small changes in microstructure can have important implications for their properties.51 This challenge was a key motivation to develop sequencecontrolled copolymerization such that the monomer sequence can be unambiguously controlled throughout the copolymerization.52 Interestingly, this issue has a different meaning in the context of supramolecular copolymers because of their dynamicity. The inter- and intra-chain exchange of monomers tends to homogenize the composition for all copolymer chains in a given system.53 Despite a growing number of strategies to achieve new supramolecular copolymer microstructures, unambiguous structural determination persists as a key challenge in supramolecular copolymerization. In this case, the monomers’ exchange and the intrinsic dynamicity complicate the characterization of the microstructure. There are several outstanding reports of supramolecular copolymers that provide strong evidence for alternate, sequenced and block microstructures; we critical review

ACS Paragon Plus Environment

Journal of the American Chemical Society 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

them here and use them as a guide to reconsider some seminal works where the microstructure determination was not possible. Alternating copolymers. Alternating supramolecular copolymers were the first category of supramolecular copolymers reported. In 1990 Lehn and co-workers presented seminal works on 1D poly(A-alt-B) structures based on multiple hydrogen-bonded complementary units.7,54 Other examples of isodesmic copolymerization using complementary units55,56,57 and organometallic supramolecular systems58 have since been reported by several groups. Another common structural motif exploits donor-acceptor monomers that copolymerize via charge transfer interactions.59-61 Evidence of copolymerization is primarily accomplished by demonstrating the inability of each monomer to homopolymerize.

Page 6 of 21

Copolymerization in water between bisurea domains functionalized with bioactive peptide or fluorescent cyanine dyes resulted in fibers with an alternated fashion as established by fluorescence experiments. Working instead in the bulk, Araki and co-workers in 2003 reported the formation of alternating supramolecular copolymers between cyclohexane tricarboxamides functionalized with linear and bulky side chains.66 The bulky monomer could not form stable homopolymers, but the authors found that it copolymerizes with the less bulky monomer in an alternating fashion based on infrared spectroscopy. An interesting example reported by Li and co-workers67 relies on hierarchical organization to form an alternating supramolecular copolymer (Figure 4b). They employ zinc porphyrins that aggregate via complementary hydrogen bonding to form the hexamer (3)6. This assembly is geometrically compatible to react with the hexadentate monomer 4. The authors use a combination of spectroscopic and static light scattering techniques to show that the monomers arrange in an alternating fashion to produce poly[(3)6-alt-4]. Periodic copolymers. Few examples of sequenced supramolecular copolymers have been reported. Würthner’s group described in 2009 the formation of poly(A2-per-B) structured copolymers via unconventional interactions between complementary monomers: a perylene bisimide (6) and an azobenzene-functionalized melamine (5, Figure 5a).68 They reported that 5 and 6 do not form long organized structures in methylcyclohexane (MCH) in a 1:1 ratio as expected for ditopic 3-folded hydrogen bonding complementary units. Contrarily, the polymerization with a 2:1 ratio of 5:6 unexpectedly formed a periodic copolymer poly[(5)2-per-6] with J-type aggregation. The authors stressed the importance of the assembly protocol to acquire organized structures, as it required an equilibration time of 15 h at 20 °C. Atomic force microscopy (AFM) shows the presence of tightly coiled helical polymers with defined pitch (6 nm) and length (100-200 nm).

Figure 4. Examples of alternating supramolecular copolymers achieved with different strategies. a) Poly(1alt-2) is primarily ordered by Coulombic interactions. b) The sequence of poly[(3)6-alt-4) is controlled by complementary hierarchical interactions. More recently, Besenius and co-workers demonstrated alternating supramolecular copolymers in aqueous solutions62,22 and by sequential growth from Au-surface.63,64 Monomers based on benzene tricarboxamides bearing positively (1) and negatively (2) charged β-sheet encoded oligopeptides form alternating copolymers poly(1-alt-2) via electrostatic interactions (Figure 4a). Another example of peptide assembly has been reported by Giuseppone’s group.65

The Würthner group more recently reported how a related distorted perylene bisimide bearing imides protected with water-soluble chains (8) copolymerizes with a planar analogue (7) to form poly[(7)4-per-82] (Figure 5b).69 The homopolymerization of 7 in water follows an isodesmic mechanism to form stiff short nanorods via π-π interaction; in contrast, 8 forms dimers that aggregate into nanoworms via hydrophobic interactions. The copolymer is obtained under kinetic control by injecting water in a tetrahydrofuran (THF) solution of dissolved 8 and 7 in a 1:2 ratio, followed by the evaporation of the THF under ambient conditions. They assign the structure through a combination of optical microscopy, nuclear magnetic resonance (NMR) spectroscopy and transmission electron microscopy (TEM). Interestingly, poly[(7)4-per-(8)2] forms long and thin flexible fibers that deviate from the morphology of the two homopolymers, evidently because the combination of the flexibility of the dimers of 8 and the rigidity of the 7-based blocks. This behavior is reminiscent of covalent elastomers where soft and hard blocks alternate to give new function.70 Block copolymers. Stitching covalent polymers together to form diblock and multi-block copolymers gives rise to properties that are more than the sum of their parts. The key to their function is their ability to microphase separate into hierarchical nanostructures, which is a function of the

ACS Paragon Plus Environment

Page 7 of 21 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Journal of the American Chemical Society

compatibility and the volume fraction ratio of the blocks. The promise of higher order structural control has motivated supramolecular chemists to create analogous supramolecular block copolymers. Again, the non-covalent nature of the bonds and the inherent difficulty to characterize such structures render this goal a great challenge. Seminal works have exploited kinetic control to retard exchange dynamics enough to furnish supramolecular block copolymers via sequential addition of monomers. In 2007, Winnik, Manners and co-workers reported the synthesis of supramolecular triblock copolymers of the form poly(A-b-Bb-A) using poly(ferrocenyldimethylsilane) (PFS) block copolymers as the supramolecular monomeric units.71 By a kinetically controlled process that they named living crystallization, they were able to create short cylindrical homomicelles that exhibit uniform growth upon the addition of more monomer.

Figure

5.

Examples

of

periodic

copolymers achieved with different strategies. a) Stable helical copolymers are obtained by unconventional hydrogen bonding interaction between complementary units of melamine-functionalized azobenzene 5 and perylene bisimide 6 resulting in poly[(5)2-per-6]. b) Planar 7 and twisted 8 perylene bisimides copolymerize under kinetic control to form long copolymers poly[(7)4-per-(8)]. They achieved the triblock architecture with control over the lengths of the blocks through the addition of a different dissolved PFS monomer. These aggregates are stable over time because of their low dynamicity. In contrast to the high dispersity observed for most supramolecular polymers produced through a cooperative mechanism, these structures exhibit low dispersity because of the “living” nature of this process, which was convincingly proven through TEM, static and dynamic light scattering (SLS, and DLS). Research on PFS-based block copolymers has since flourished. Recent examples include the formation of regular crossed co-micelles,72 di- and tri-block co-micelles,73 fluorescent multi-block micelles,74,75 co-micelles between linear and brush units,76 photo-responsive systems,77 and multidimensional micelles.78-80 Other types of monomer have been found to also undergo living crystallization, including poly(p-phenylvinylene)s81 and poly(di-n-hexylfluorene)s.82 The latter, based on electron poor (9) and electron rich (10) poly(di-n-hexylfluorene)s, copolymerizes to form poly(10-b9-b-10), which exhibits long range exciton transport (Figure 6a).

supramolecular

ACS Paragon Plus Environment

Journal of the American Chemical Society 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 8 of 21

Figure 6. Examples of supramolecular block copolymers. a) Covalent block copolymers 9 and 10 feature a crystalline midblock of poly(di-n-hexylfluorene) and are used as monomers to obtain triblock micelles poly(9-b-10-b-9) via seeded multistep addition and crystallization driven copolymerization. b) Supramolecular copolymerization of zinc porphyrin 11 followed by copper porphyrin 12 under kinetic control yields a triblock copolymer poly(11-b-12-b-11). c) Supramolecular block copolymers based on electron rich and electron poor hexabenzocorone-base monomers 13 and 14, respectively, are obtained by a multistep synthesis. c) Supramolecular multiblock copolymers are obtained between tripyridylamine tricarboxamide 15 and triphenylamine tricarboxamide 16 under thermodynamic control. Direct mixing of poly(16) to poly(15) or copolymerization from a mixture of monomers lead to the same copolymer microstructure poly[(15)-co-(14)]. Pavan, Takeuchi, Sugiyasu, and co-workers83 recently demonstrated a seeded-growth copolymerization strategy to create kinetically stable supramolecular block copolymers. They generate seeds by sonicating homopolymers of zincbased porphyrin 11, and these seeds can be used to again grow poly(11) by adding additional 11 or to produce poly(12b-11-b-12) by adding the copper porphyrin 12 (Figure 6b). This process yields aggregates with a predictable length based on the monomer/seed ratio along with a narrow polydispersity. Remarkably, the triblock copolymer exhibits greater stability toward the presence of a denaturant compared to the poly(11) homopolymer. Molecular dynamics simulations reveal that monomer exchange only occurs at the end groups of the polymer, suggesting that the more

stable poly(12) end blocks compartmentalize and protect the poly(11) mid-block. Another exceptional example is presented by Aida’s group, 84 who reported the block copolymerization of hexabenzacoronenes based-monomers through sequential steps (Figure 6c). The bipyridine-terminated hexabenzacoronenes units (13) homopolymerize in methanol to give bundles of poly(13). Metallation of the pendant bipyridine units with copper results in a dispersion of charged assemblies poly(13*). These aggregates are then dispersed in acetone and fragmented by sonication to create seeds. Upon adding 14 and cooling from 50 °C to 25 °C, the seeds initiate the supramolecular polymerization of 14 to give a mixture of diblock and triblock copolymers, poly(13*-b-14) and poly(14-b-13*-b-14), respectively.

ACS Paragon Plus Environment

Page 9 of 21 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Journal of the American Chemical Society

Similarly to the synthesis of covalent block copolymers, the majority of supramolecular block copolymers are produced via a kinetic approach in which supramolecular polymers with low dynamicity can be fed with different monomers to form block structures. Their thermodynamic stability is in most cases limited, though. In contrast to these reports, we prepared supramolecular block copolymers under thermodynamic control with tripyridylamine tricarboxamide 15 and triphenylamine tricarboxamide 16. We found that supramolecular multiblock copolymers poly(15co-16) are produced both by directly mixing the homopolymers poly(15) and poly(16) and also by copolymerization of the molecularly dissolved monomers via a cooling procedure (Figure 6d).50 By combining optical spectroscopy and microscopy with the theoretical model developed for two-component supramolecular polymerization, we were able to identify the microstructure of the polymer, finding that the block length distribution is closely related to the enthalpy of co-interaction of the different monomers with respect to the enthalpy of the related homo-interaction. Random and unresolved structures. Some of the earliest studies in the field of one-dimensional supramolecular polymerization consist of copolymerization experiments. Inspired by the work of Green and co-workers85–87 on helical covalent polymers, these experiments were designed to probe the amplification of chirality in helical onedimensional supramolecular polymers. Typical experiments include mixing enantiomers of a monomer together (majority rules experiment),88–92 mixing achiral monomers with chiral enantiopure analogues (sergeants and soldiers experiment),90,93–103 and mixing chiral enantiopure monomers in different ratios with an achiral counterpart (diluted majority rules experiment).104–106 The monomers in these cases typically comprise identical cores and only differ in the structure of their side chains. The mixtures were predominantly characterized using spectroscopic measurements such as fluorescence, ultraviolet-visible (UVVis) spectroscopy and especially circular dichroism (CD). Apart from some notable exceptions,107 very few studies have corroborated the spectroscopic results with other types of characterization techniques. Usually, a stochastic incorporation of the monomers during the copolymerization process is assumed, but it has been difficult to experimentally assess if monomer incorporation is indeed fully random. Interestingly, several supramolecular copolymer systems exhibit a linear response in optical activity as a function of e.e. (majority rules) or percentage of chiral sergeant added (sergeant and soldiers). This lack of chiral amplification is a useful indicator for self-sorting (Figure 2), as it points to helical supramolecular homopolymers co-existing rather than mixing into copolymers. Detailed investigations by the groups of Würthner, Sanchez and our group suggest that these systems are often kinetically trapped, requiring either heat97,108,109 or the addition of a co-solvent110 to initiate monomer exchange. More recently, unresolved copolymer structures have been reported with monomers that have different supramolecular units96 or have different cores.92,96,111 For example, the copolymerization of a chiral benzene tricarboxamide with an achiral benzene trisurea forms a kinetically trapped copolymer that spontaneously evolves

into self-sorted homopolymers after 1 h.96 In contrast, the copolymerization of N-centered and CO-centered benzene tricarboxamides most likely results in an alternating copolymer based on their donor-acceptor structure and chiral amplification studies.92 Finally, mixtures of copolymers of oligo(p-phenylvinylene) with different sizes and opposite chirality assemble into homopolymers with either preferred P-helicity or M-helicity at room temperature. Heating followed by slow cooling induces the formation of copolymers with the helical sense bias depending on the composition of the mixture.111 In all of these cases, we rely on chemical intuition to propose their microstructures, but current analytical tools are insufficient to unambiguously assign their structures. In some cases, the steric bulk of a monomer’s side chains is too demanding, precluding it from mixing with another monomer to form a copolymer.104,112 However, our group recently reported that a dendronized-pegylated benzene tricarboxamide is too bulky to homopolymerize but forms stable aggregates when mixed with a linearly pegylated benzene tricarboxamide in water.113 A 2:1 ratio of dendronized:linear monomer exhibits a higher degree of order, slower dynamics, and higher stability compared to the homopolymer of the linearly pegylated monomer based on UV-Vis spectroscopy, hydrogen-deuterium exchange experiments, and cryoTEM imaging. Although an experimental elucidation of its structure remains elusive, molecular dynamic simulations suggest that a periodic 2:1 copolymer composition allows for an optimal balance between steric hindrance of the dendronized side chains and the stability afforded by intermolecular hydrogen bonds. Another elegant system based on bipyridine-derived benzene tricarboxamides has been described by Brunsveld and co-workers as a functional material in biological systems.114,115 Monomers with charged groups, mannose groups, fluorescent dyes, and biotin residues combine to form supramolecular copolymers in aqueous media, and the ratio of these components dictates their efficacy in binding and detecting bacteria. However, detailed information on the sequence and actual incorporation of the different monomers remains very challenging to obtain.

 CHARACTERIZATION CHALLENGES The growing interest in developing non-covalent copolymers requires parallel progress and standardization of characterization techniques. As discussed above, such uniformity remains elusive. Other established fields, such as covalent copolymers, teach us that having a systematic, universally applied characterization routine is crucial to achieve concrete progress. However, the dynamicity of supramolecular polymers and the non-covalent nature of the bonds between the monomers render microstructural analyses extremely challenging. Nevertheless, nowadays a clear trend on the supramolecular homopolymers characterization routine has emerged. The prevalent strategy combines spectroscopic measurements (especially useful for variable temperature or variable solvent measurements), microscopy (the most used are AFM, TEM, cryoTEM, and super resolution fluorescence microcopy), mass spectrometry, and light scattering techniques. These experimental techniques are complemented with molecular 9

ACS Paragon Plus Environment

Journal of the American Chemical Society 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

dynamics simulations and thermodynamic mass-balance models. The outstanding reports on supramolecular copolymers in which a detailed structure analysis is reported50,84,116 adapt and combine these techniques to resolve the supramolecular copolymer microstructure. Spectroscopy measurements (UV, CD, fluorescence, etc.) are powerful tools to unravel both the copolymerization mechanism50 and composition117 of a given system by quantifying changes in spectroscopic features. Thereafter, spectroscopic results can be used in combination with theoretical modeling to predict the microstructure and the relative stability of the copolymer with respect to the homopolymers.50,117 The theoretical microstructural analysis can then be corroborated with judicious use of microscopy techniques. If applicable, TEM has proven to be an excellent tool to unravel structure.84 Otherwise, recent developments in super resolution fluorescence microscopy demonstrate that it can be helpful in unravelling polymer growth118, dynamics53,119 and structure.120-122 Furthermore, hydrogendeuterium exchange followed by mass spectrometry is a valuable tool for following the dynamics of supramolecular systems in water.123,113 What is striking from these seminal works is the lack of a leading experimental technique to unambiguously determine a copolymer’s structure and the consequential need to bridge this gap with a combination of multiple techniques. This situation differs from that of covalent copolymers, where 1H and 13C NMR spectroscopy are the primary tools for microstructural determination. We believe that further studies for developing techniques which allow similar analyses for supramolecular analogues are needed and will permit more efficient growth in the field. Microscopy, modeling, and molecular dynamics simulations have so far resulted in promising strategies. More progress in these directions could lead towards the possibility of analyzing in detail the subtle structural details of supramolecular copolymers.

 APPLICATIONS AND FUTURE PERSPECTIVES The development of covalent copolymers has proven to be invaluable with time. Covalent copolymerization has opened the door to creating materials with properties simply not achievable with homopolymerization or blending. Considering supramolecular copolymers in this framework, we can dream of the advancements achievable through mastering dynamic multicomponent structures. Their properties are particularly well suited for managing complexity through adaptability. For biological applications, the prospect of including multiple units—such as sensors, bio-active molecules, and catalysts—in a defined order in non-covalent systems would be a breakthrough for synthetic bio-systems that require cooperative feedback between multiple systems.124,125,114 The same effects can be exploited in organic electronics where tuning the arrangements of dynamic copolymers would optimize the optoelectronic properties of, for example, photoelectronic switches, sensors and chiral devices.126-128 This perspective is meant to initiate a cooperative effort to advance this new and promising field. Despite the great progress achieved in the last years, much more must come. We believe that the field of supramolecular copolymerization needs comprehensive growth spanning

Page 10 of 21

strategic molecular design to the exploitation of theoretical models to the development of powerful characterization techniques. Achievable through collaboration and standardized analytical routines, these efforts will bring supramolecular copolymers to be essential in cutting edge technology markets.

 ACKNOWLEDGEMENTS The authors like to thank Dr. H.M.M. ten Eikelder and Dr. A.J. Markvoort for fruitful discussions and the mathematical models describing the supramolecular copolymerizations. We acknowledge funding from the European Union's Horizon 2020 research and innovation program under grant agreement No 705701 (N.J.V.Z.), the Dutch Science Foundation (NWO ECHO Grant No. 712.016.003, L.N.J.d.W.), and the Dutch Ministry of Education, Culture and Science (Gravity program 024.001.035).  AUTHOR INFORMATION Corresponding Author: [email protected] ORCID Beatrice Adelizzi ORCID 0000-0002-7763-8999 Nathan Van Zee ORCID 0000-0002-7899-9163 Anja Palmans ORCID 0000-0002-7201-1548 E.W. Meijer ORCID 0000-0003-4126-7492

 REFERENCES (1)

De Greef, T. F. A.; Smulders, M. M. J.; Wolffs, M.; Schenning, A. P. H. J.; Sijbesma, R. P.; Meijer, E. W. Supramolecular Polymerization. Chem. Rev. 2009, 109 (11), 5687–5754.

(2)

Liu, M.; Zhang, L.; Wang, T.; Yashima, E.; Maeda, K.; Iida, H.; Furusho, Y.; Nagai, K. Supramolecular Chirality in SelfAssembled Systems. Chem. Rev. 2015, 115 (15), 6102–6211.

(3)

Yashima, E.; Ousaka, N.; Taura, D.; Shimomura, K.; Ikai, T.; Maeda, K. Supramolecular Helical Systems: Helical Assemblies of Small Molecules, Foldamers, and Polymers with Chiral Amplification and Their Functions. Chem. Rev. 2016, 116 (22), 13752–13990.

(4)

Das, A.; Ghosh, S. H-Bonding Directed Programmed Supramolecular Assembly of Naphthalene-diimide (NDI) Derivatives. Chem. Commun. 2016, 52 (42), 6860–6872.

(5)

Yang, S. K.; Zimmerman, S. C. Hydrogen Bonding Modules for Use in Supramolecular Polymers. Isr. J. Chem. 2013, 53 (8), 511–520.

(6)

van der Zwaag, D.; de Greef, T. F. A.; Meijer, E. W. Programmable Supramolecular Polymerizations. Angew. Chem. Int. Ed. 2015, 54 (29), 8334–8336.

(7)

Fouquey, C.; Lehn, J.-M.; Levelut, A.-M. Molecular Recognition Directed Self-assembly of Supramolecular Liquid Crystalline Polymers from Complementary Chiral Components. Adv. Mater. 1990, 2 (5), 254–257.

(8)

Odian, G. G. Principle of Polymerization, Third Ed.; Wiley, Ed.; 1991.

(9)

Szwarc, M. Molecular Weight Distribution of “Living Polymers,” II. Effect of Impurities. J. Phys. Chem. 1958, 62 (5), 568–569.

(10)

Hirschberg, J. H. K. K.; Beijer, F. H.; Van Aert, H. A.;

10

ACS Paragon Plus Environment

Page 11 of 21 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Journal of the American Chemical Society Magusin, P. C. M. M.; Sijbesma, R. P.; Meijer, E. W. Supramolecular Polymers from Linear Telechelic Siloxanes with Quadruple-Hydrogen-Bonded Units. Macromolecules 1999, 32 (8), 2696–2705.

(11)

Xu, Q.; Han, B.; Yan, H. Effect of Cosolvents on the Precipitation Polymerization of Acrylic Acid in Supercritical Carbon Dioxide. Polymer 2001, 42 (4), 1369–1373.

(12)

Korevaar, P. A.; Schaefer, C.; de Greef, T. F. A.; Meijer, E. W. Controlling Chemical Self-Assembly by Solvent-Dependent Dynamics. J. Am. Chem. Soc. 2012, 134 (32), 13482–13491.

(13)

(14)

Weng, W.; Li, Z.; Jamieson, A. M.; Rowan, S. J. Control of Gel Morphology and Properties of a Class of MetalloSupramolecular Polymers by Good/Poor Solvent Environments. Macromolecules 2009, 42 (1), 236–246. Greenley, R. Z. Free Radical Copolymerization Reactivity Ratios. The Wiley Database of Polymer Properties. April 2003.

(15)

Flory, P. J. Principles of Polymer Chemistry; Cornell University press, 1953.

(16)

Burnworth, M.; Tang, L.; Kumpfer, J. R.; Duncan, A. J.; Beyer, F. L.; Fiore, G. L.; Rowan, S. J.; Weder, C. Optically Healable Supramolecular Polymers. Nature 2011, 472, 334– 337.

(17)

Mark, J.; Ngai, K.; Graessley, W.; Mandelkern, L.; Samulski, E.; Koenig, J.; Wignall, G. Physical Properties of Polymers, 3rd ed.; Cambridge University Press: Cambridge, 2004.

(18)

Verbiest, T.; Elshocht, S. Van; Kauranen, M.; Hellemans, L.; Snauwaert, J.; Nuckolls, C.; Katz, T. J.; Persoons, A. Strong Enhancement of Nonlinear Optical Properties Through Supramolecular Chirality. Science 1998, 282 (5390), 913–915.

(19)

Jelley, E. E. Spectral Absorption and Fluorescence of Dyes in the Molecular State. Nature 1936, 138, 1009.

(20)

Eisfeld, A.; Briggs, J. S. The J- and H-Bands of Organic Dye Aggregates. Chem. Phys. 2006, 324 (2–3), 376–384.

(21)

Lortie, F.; Boileau, S.; Bouteiller, L.; Chassenieux, C.; Lauprêtre, F. Chain Stopper-Assisted Characterization of Supramolecular Polymers. Macromolecules 2005, 38 (12), 5283–5287.

(28)

Ogi, S.; Sugiyasu, K.; Manna, S.; Samitsu, S.; Takeuchi, M. Living Supramolecular Polymerization Realized through a Biomimetic Approach. Nat. Chem. 2014, 6 (3), 188–195.

(29)

Endo, M.; Fukui, T.; Jung, S. H.; Yagai, S.; Takeuchi, M.; Sugiyasu, K. Photoregulated Living Supramolecular Polymerization Established by Combining Energy Landscapes of Photoisomerization and NucleationElongation Processes. J. Am. Chem. Soc. 2016, 138 (43), 14347–14353.

(30)

Aliprandi, A.; Mauro, M.; De Cola, L. Controlling and Imaging Biomimetic Self-Assembly. Nat. Chem. 2016, 8 (1), 10–15.

(31)

Korevaar, P. A.; George, S. J.; Markvoort, A. J.; Smulders, M. M. J.; Hilbers, P. A. J.; Schenning, A. P. H. J.; De Greef, T. F. A.; Meijer, E. W. Pathway Complexity in Supramolecular Polymerization. Nature 2012, 481 (7382), 492–496.

(32)

Korevaar, P. A.; de Greef, T. F. A.; Meijer, E. W. Pathway Complexity in π-Conjugated Materials. Chem. Mater. 2014, 26 (1), 576–586.

(33)

van der Zwaag, D.; Pieters, P. A.; Korevaar, P. A.; Markvoort, A. J.; Spiering, A. J. H.; de Greef, T. F. A.; Meijer, E. W. Kinetic Analysis as a Tool to Distinguish Pathway Complexity in Molecular Assembly: An Unexpected Outcome of Structures in Competition. J. Am. Chem. Soc. 2015, 137 (39), 12677–12688.

(34)

Valera, J. S.; Gómez, R.; Sánchez, L. Tunable Energy Landscapes to Control Pathway Complexity in SelfAssembled N-Heterotriangulenes: Living and Seeded Supramolecular Polymerization. Small 2018, 14 (3), 1–9.

(35)

Ogi, S.; Fukui, T.; Jue, M. L.; Takeuchi, M.; Sugiyasu, K. Kinetic Control over Pathway Complexity in Supramolecular Polymerization through Modulating the Energy Landscape by Rational Molecular Design. Angew. Chem. Int. Ed. 2014, 53 (52), 14363–14367.

(36)

Liu, M.; Zhang, L.; Wang, T. Supramolecular Chirality in Self-Assembled Systems. Chem. Rev. 2015, 115 (15), 7304– 7397.

(37)

Biedermann, F.; Schneider, H. J. Experimental Binding Energies in Supramolecular Complexes. Chem. Rev. 2016, 116 (9), 5216–5300.

(38)

Venkata Rao, K.; Miyajima, D.; Nihonyanagi, A.; Aida, T. Thermally Bisignate Supramolecular Polymerization. Nat. Chem. 2017, 9, 1133–1139.

(22)

Ahlers, P.; Fischer, K.; Spitzer, D.; Besenius, P. Dynamic Light Scattering Investigation of the Kinetics and Fidelity of Supramolecular Copolymerizations in Water. Macromolecules 2017, 50 (19), 7712–7720.

(23)

Smulders, M. M. J.; Nieuwenhuizen, M. M. L.; Grossman, M.; Filot, I. A. W.; Lee, C. C.; De Greef, T. F. A.; Schenning, A. P. H. J.; Palmans, A. R. A.; Meijer, E. W. Cooperative Two-Component Self-Assembly of Mono- and Ditopic Monomers. Macromolecules 2011, 44 (16), 6581–6587.

(39)

Kulkarni, C.; Korevaar, P. A.; Bejagam, K. K.; Palmans, A. R. A.; Meijer, E. W.; George, S. J. Solvent Clathrate Driven Dynamic Stereomutation of a Supramolecular Polymer with Molecular Pockets. J. Am. Chem. Soc. 2017, 139 (39), 13867– 13875.

(24)

Würthner, F. Supramolecular Polymerization: Living It up. Nat. Chem. 2014, 6 (3), 171–173.

(40)

(25)

Kang, J.; Miyajima, D.; Mori, T.; Inoue, Y.; Itoh, Y.; Aida, T. A Rational Strategy for the Realization of Chain-Growth Supramolecular Polymerization. Science 2015, 347 (6222), 646–651.

Cantekin, S.; Nakano, Y.; Everts, J. C.; van der Schoot, P.; Meijer, E. W.; Palmans, A. R. A. A Stereoselectively Deuterated Supramolecular Motif to Probe the Role of Solvent during Self-Assembly Processes. Chem. Commun. 2012, 48 (32), 3803–3805.

(41)

Ogi, S.; Stepanenko, V.; Thein, J.; Würthner, F. Impact of Alkyl Spacer Length on Aggregation Pathways in Kinetically Controlled Supramolecular Polymerization. J. Am. Chem. Soc. 2016, 138 (2), 670–678.

Dong, S.; Leng, J.; Feng, Y.; Liu, M.; Stackhouse, C. J.; Schönhals, A.; Chiappisi, L.; Gao, L.; Chen, W.; Shang, J.; et al. Structural Water as an Essential Comonomer in Supramolecular Polymerization. Sci. Adv. 2017, 3 (11), 1–9.

(42)

Ogi, S.; Stepanenko, V.; Würthner, F. Mechanism Seeded Supramolecular Bisimide Organogelator. J. 3300–3307.

Ortony, J. H.; Qiao, B.; Newcomb, C. J.; Keller, T. J.; Palmer, L. C.; Deiss-Yehiely, E.; Olvera De La Cruz, M.; Han, S.; Stupp, S. I. Water Dynamics from the Surface to the Interior of a Supramolecular Nanostructure. J. Am. Chem. Soc. 2017, 139 (26), 8915–8921.

(43)

Johnson, R. S.; Yamazaki, T.; Kovalenko, A.; Fenniri, H.

(26)

(27)

Sugiyasu, K.; Takeuchi, M.; of Self-Assembly Process and Polymerization of Perylene Am. Chem. Soc. 2015, 137 (9),

11

ACS Paragon Plus Environment

Journal of the American Chemical Society 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Molecular Basis for Water-Promoted Supramolecular Chirality Inversion in Helical Rosette Nanotubes. J. Am. Chem. Soc. 2007, 129 (17), 5735–5743.

Page 12 of 21

(59)

Das, A.; Ghosh, S. Supramolecular Assemblies by ChargeTransfer Interactions between Donor and Acceptor Chromophores. Angew. Chemie Int. Ed. 2014, 53 (8), 2038– 2054.

(60)

Jalani, K.; Kumar, M.; George, S. J. Mixed Donor-Acceptor Charge-Transfer Stacks Formed via Hierarchical SelfAssembly of a Non-Covalent Amphiphilic Foldamer. Chem. Commun. 2013, 49 (45), 5174–5176.

(44)

Gillissen, M. A. J.; Koenigs, M. M. E.; Spiering, J. J. H.; Vekemans, J. A. J. M.; Palmans, A. R. A.; Voets, I. K.; Meijer, E. W. Triple Helix Formation in Amphiphilic Discotics: Demystifying Solvent Effects in Supramolecular SelfAssembly. J. Am. Chem. Soc. 2014, 136 (1), 336–343.

(45)

Liu, C.; Jin, Q.; Lv, K.; Zhang, L.; Liu, M. Water Tuned the Helical Nanostructures and Supramolecular Chirality in Organogels. Chem. Commun. 2014, 50 (28), 3702–3705.

(61)

Rizkov, D.; Gun, J.; Lev, O.; Sicsic, R.; Melman, A. DonorAcceptor-Promoted Gelation of Polyaromatic Compounds. Langmuir 2005, 21 (26), 12130–12138.

(46)

Liu, Y.; Jia, Y.; Zhu, E.; Liu, L.; Qiao, Y.; Che, G.; Yin, B. Supramolecular Helical Nanofibers Formed by an Achiral Monopyrrolotetrathiafulvalene Derivative: Water-Triggered Gelation and Chiral Evolution. New J. Chem. 2017, 41 (19), 11060–11068.

(62)

Frisch, H.; Unsleber, J. P.; Lüdeker, D.; Peterlechner, M.; Brunklaus, G.; Waller, M.; Besenius, P. pH-Switchable Ampholytic Supramolecular Copolymers. Angew. Chem. Int. Ed. 2013, 52 (38), 10097–10101.

(47)

Yu, X.; Xie, D.; Lan, H.; Li, Y.; Zhen, X.; Ren, J.; Yi, T. Effect of Water on the Supramolecular Assembly and Functionality of a Naphthalimide Derivative: Tunable Honeycomb Structure with Mechanochromic Properties. J. Mater. Chem. C 2017, 5 (24), 5910–5916.

(63)

Frisch, H.; Fritz, E. C.; Stricker, F.; Schmüser, L.; Spitzer, D.; Weidner, T.; Ravoo, B. J.; Besenius, P. Kinetically Controlled Sequential Growth of Surface-Grafted Chiral Supramolecular Copolymers. Angew. Chem. Int. Ed. 2016, 55 (25), 7242–7246.

(48)

Zee, N. J. Van; Adelizzi, B.; Mabesoone, M. F. J.; Meng, X.; Aloi, A.; Zha, H.; Lutz, M.; Filot, I. A. W.; Palmans, A. R. A.; Meijer, E. W. Potential Enthalpic Energy of Water in Oils Exploited to Control Supramolecular Structure. Nature 2018, 558 (7708), 100–103.

(64)

Engel, S.; Spitzer, D.; Rodrigues, L. L.; Fritz, E. C.; Straßburger, D.; Schönhoff, M.; Ravoo, B. J.; Besenius, P. Kinetic Control in the Temperature-Dependent Sequential Growth of Surface-Confined Supramolecular Copolymers. Faraday Discuss. 2017, 204, 53–67.

(49)

Xing, P.; Li, Y.; Wang, Y.; Li, P. Z.; Chen, H.; Phua, S. Z. F.; Zhao, Y. Water-Binding-Mediated Gelation/Crystallization and Thermosensitive Superchirality. Angew. Chem. Int. Ed. 2018, 57 (26), 7774–7779.

(65)

Xiang, Y.; Moulin, E.; Buhler, E.; Maaloum, M.; Fuks, G.; Giuseppone, N. Hydrogen-Bonded Multifunctional Supramolecular Copolymers in Water. Langmuir 2015, 31 (28), 7738–7748.

(50)

Adelizzi, B.; Aloi, A.; Markvoort, A. J.; Ten Eikelder, H. M. M.; Voets, I. K.; Palmans, A. R. A.; Meijer, E. W. Supramolecular Block Copolymers under Thermodynamic Control. J. Am. Chem. Soc. 2018, 140 (23), 7168–7175.

(66)

Takasawa, R.; Murota, K.; Yoshikawa, I.; Araki, K. StericFactor-Directed Alternating Supramolecular Copolymer Composed of Hydrogen-Bonded Cyclohexanetricarboxamide Units. Macromol. Rapid Commun. 2003, 24 (4), 335–339.

(51)

Ouchi, M.; Nakano, M.; Nakanishi, T.; Sawamoto, M. Alternating Sequence Control for Carboxylic Acid and Hydroxy Pendant Groups by Controlled Radical Cyclopolymerization of a Divinyl Monomer Carrying a Cleavable Spacer. Angew. Chem. Int. Ed. 2016, 55 (47), 14584–14589.

(67)

Chen, S. G.; Yu, Y.; Zhao, X.; Ma, Y.; Jiang, X. K.; Li, Z. T. Highly Stable Chiral (A)6-B Supramolecular Copolymers: A Multivalency-Based Self-Assembly Process. J. Am. Chem. Soc. 2011, 133 (29), 11124–11127.

(68)

Yagai, S.; Hamamura, S.; Wang, H.; Stepanenko, V.; Seki, T.; Unoike, K.; Kikkawa, Y.; Karatsu, T.; Kitamura, A.; Würthner, F. Unconventional Hydrogen-Bond-Directed Hierarchical Co-Assembly between Perylene Bisimide and Azobenzene-Functionalized Melamine. Org. Biomol. Chem. 2009, 7 (19), 3926–3929.

(69)

Görl, D.; Zhang, X.; Stepanenko, V.; Würthner, F. Supramolecular Block Copolymers by Kinetically Controlled Co-Self-Assembly of Planar and Core-Twisted Perylene Bisimides. Nat. Commun. 2015, 6, 7009.

(70)

Hepburn, C. Polyurethane Netherlands: Dordrecht, 1992.

(71)

Wang, X.; Guerin, G.; Wang, H.; Wang, Y.; Manners, I.; Winnik, M. a. Cylindrical Block Copolymer Micelles and Co-Micelles of Controlled Length and Architecture. Science 2007, 317 (5838), 644–647.

(72)

Qiu, H.; Russo, G.; Rupar, P. A.; Chabanne, L.; Winnik, M. A.; Manners, I. Tunable Supermicelle Architectures from the Hierarchical Self-Assembly of Amphiphilic Cylindrical B-A-B Triblock Co-Micelles. Angew. Chem. Int. Ed. 2012, 51 (47), 11882–11885.

(73)

Rupar, P. A.; Chabanne, L.; Winnik, M. A.; Manners, I. NonCentrosymmetric Cylindrical Micelles by Unidirectional Growth. Science 2012, 337 (6094), 559–562.

(74)

Hudson, Z. M.; Lunn, D. J.; Winnik, M. A.; Manners, I.

(52)

Lutz, J.-F.; Ouchi, M.; Liu, D. R.; Sawamoto, M. SequenceControlled Polymers. Science 2013, 341 (6146), 1238149– 1238149.

(53)

Albertazzi, L.; Hofstad, R. W. v. d.; Meijer, E. W. Probing Exchange Pathways in Super-Resolution Microscopy. Science 2014, 491 (6183), 10–15.

(54)

Lehn, J.-M. Perspectives in Supramolecular Chemistry From Molecular Recognition towards Molecular Information Processing and Self-Organization. Angew. Chem. Int. Ed. 1990, 29 (11), 1304–1319.

(55)

Sivakova, S.; Rowan, S. J. Nucleobases as Supramolecular Motifs. Chem. Soc. Rev. 2005, 34 (1), 9–21.

(56)

Sijbesma, R. P.; Beijer, F. H.; Brunsveld, L.; Folmer, B. J. B.; Hirschberg, J. H. K. K.; Lange, R. F. M.; Lowe, J. K. L.; Meijer, E. W. Reversible Polymers Formed from SelfComplementary Monomers Using Quadruple Hydrogen Bonding. Science 1997, 278 (5343), 1601–1604.

(57)

(58)

Blight, B. A.; Camara-Campos, A.; Djurdjevic, S.; Kaller, M.; Leigh, D. A.; McMillan, F. M.; McNab, H.; Slawin, A. M. Z. AAA−DDD Triple Hydrogen Bond Complexes. J. Am. Chem. Soc. 2009, 131 (39), 14116–14122. Gohy, J.-F. Metallo-Supramolecular Block Copolymer Micelles. Coord. Chem. Rev. 2009, 253 (17), 2214–2225.

Elastomers;

Springer

12

ACS Paragon Plus Environment

Page 13 of 21 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Journal of the American Chemical Society Colour-Tunable Fluorescent Multiblock Micelles. Nat. Commun. 2014, 5, 1–8.

(75)

Xu, J.; Zhou, H.; Yu, Q.; Manners, I.; Winnik, M. A. Competitive Self-Assembly Kinetics as a Route to Control the Morphology of Core-Crystalline Cylindrical Micelles. J. Am. Chem. Soc. 2018, 140 (7), 2619–2628.

(76)

Finnegan, J. R.; Lunn, D. J.; Gould, O. E. C.; Hudson, Z. M.; Whittell, G. R.; Winnik, M. A.; Manners, I. Gradient Crystallization-Driven Self-Assembly: Cylindrical Micelles With “patchy” segmented Coronas via the Coassembly of Linear and Brush Block Copolymers. J. Am. Chem. Soc. 2014, 136 (39), 13835–13844.

(77)

Gao, Y.; Qiu, H.; Zhou, H.; Li, X.; Harniman, R.; Winnik, M. A.; Manners, I. Crystallization-Driven Solution SelfAssembly of Block Copolymers with a Photocleavable Junction. J. Am. Chem. Soc. 2015, 137 (6), 2203–2206.

(78)

Li, X.; Gao, Y.; Boott, C. E.; Winnik, M. A.; Manners, I. NonCovalent Synthesis of Supermicelles with Complex Architectures Using Spatially Confined Hydrogen-Bonding Interactions. Nat. Commun. 2015, 6, 1–8.

(79)

Li, X.; Gao, Y.; Boott, C. E.; Hayward, D. W.; Harniman, R.; Whittell, G. R.; Richardson, R. M.; Winnik, M. A.; Manners, I. “Cross” Supermicelles via the Hierarchical Assembly of Amphiphilic Cylindrical Triblock Comicelles. J. Am. Chem. Soc. 2016, 138 (12), 4087–4095.

Chem. Soc. 2005, 127 (15), 5490–5494. (89)

Jin, W.; Fukushima, T.; Niki, M.; Kosaka, A.; Ishii, N.; Aida, T. Self-Assembled Graphitic Nanotubes with One-Handed Helical Arrays of a Chiral Amphiphilic Molecular Graphene. Proc. Natl. Acad. Sci. 2005, 102 (31), 10801–10806.

(90)

Langeveld-Voss, B. M. W.; Waterval, R. J. M.; Janssen, R. A. J.; Meijer, E. W. Principles of “Majority Rules” and “Sergeants and Soldiers” Applied to the Aggregation of Optically Active Polythiophenes: Evidence for a Multichain Phenomenon. Macromolecules 1999, 32 (1), 227–230.

(91)

Isare, B.; Linares, M.; Zargarian, L.; Fermandjian, S.; Miura, M.; Motohashi, S.; Vanthuyne, N.; Lazzaroni, R.; Bouteiller, L. Chirality in Dynamic Supramolecular Nanotubes Induced by a Chiral Solvent. Chem. Eur. J. 2010, 16 (1), 173– 177.

(92)

Stals, P. J. M.; Everts, J. C.; de Bruijn, R.; Filot, I. A. W.; Smulders, M. M. J.; Martín-Rapún, R.; Pidko, E. A.; de Greef, T. F. A.; Palmans, A. R. A.; Meijer, E. W. Dynamic Supramolecular Polymers Based on Benzene-1,3,5Tricarboxamides: The Influence of Amide Connectivity on Aggregate Stability and Amplification of Chirality. Chem. Eur. J. 2010, 16 (3), 810–821.

(93)

Palmans, A. R. A.; Vekemans, J. a. J. M.; Havinga, E. E.; Meijer, E. W. Sergeants-and-Soldiers Principle in Chiral Columnar Stacks of Disc-Shaped Molecules with C3 Symmetry. Angew. Chem. Int. Ed. 1997, 36 (23), 2648–2651.

(94)

Kim, T.; Mori, T.; Aida, T.; Miyajima, D. Dynamic Propeller Conformation for the Unprecedentedly High Degree of Chiral Amplification of Supramolecular Helices. Chem. Sci. 2016, 7 (11), 6689–6694.

(95)

Smulders, M. M. J.; Schenning, A. P. H. J.; Meijer, E. W. Insight into the Mechanisms of Cooperative Self-Assembly: The “Sergeants-and-Soldiers” Principle of Chiral and Achiral C3-Symmetrical Discotic Triamides. J. Am. Chem. Soc. 2008, 130 (2), 606–611.

(96)

van Gorp, J. J.; Vekemans, J. A. J. M.; Meijer, E. W. C 3 Symmetrical Supramolecular Architectures: Fibers and Organic Gels from Discotic Trisamides and Trisureas. J. Am. Chem. Soc. 2002, 124 (49), 14759–14769.

(80)

Hudson, Z. M.; Boott, C. E.; Robinson, M. E.; Rupar, P. A.; Winnik, M. A.; Manners, I. Tailored Hierarchical Micelle Architectures Using Living Crystallization-Driven SelfAssembly in Two Dimensions. Nat. Chem. 2014, 6 (10), 893– 898.

(81)

Shin, S.; Menk, F.; Kim, Y.; Lim, J.; Char, K.; Zentel, R.; Choi, T. L. Living Light-Induced Crystallization-Driven Self-Assembly for Rapid Preparation of Semiconducting Nanofibers. J. Am. Chem. Soc. 2018, 140 (19), 6088–6094.

(82)

Jin, X.; Price, M. B.; Finnegan, J. R.; Boott, C. E.; Richter, J. M.; Rao, A.; Menke, S. M.; Friend, R. H.; Whittell, G. R.; Manners, I. Long-Range Exciton Transport in Conjugated Polymer Nanofibers Prepared by Seeded Growth. Science 2018, 360 (6391), 897–900.

(83)

Sugiyasu, K.; Jung, S. H.; Bochicchio, D.; Pavan, G. M.; Takeuchi, M.; Sugiyasu, K. A Block Supramolecular Polymer and Its Kinetically Enhanced Stability. J. Am. Chem. Soc. 2018, 140 (33), 10570–10577.

(97)

Lohr, A.; Würthner, F. Time-Dependent Amplification of Helical Bias in Self-Assembled Dye Nanorods Directed by the Sergeants-and-Soldiers Principle. Chem. Commun. 2008, 19, 2227.

(84)

Zhang, W.; Jin, W.; Fukushima, T.; Saeki, A.; Seki, S.; Aida, T. Supramolecular Linear Heterojunction Composed of Graphite-like Semiconducting Nanotubular Segments. Science 2011, 334 (6054), 340–343.

(98)

Ajayaghosh, A.; Varghese, R.; George, S. J.; Vijayakumar, C. Transcription and Amplification of Molecular Chirality to Oppositely Biased Supramolecular π Helices. Angew. Chem. Int. Ed. 2006, 45 (7), 1141–1144.

(85)

Green, M. M.; Andreola, C.; Muñoz, B.; Reidy, M. P.; Zero, K. Macromolecular Stereochemistry: A Cooperative Deuterium Isotope Effect Leading to a Large Optical Rotation. J. Am. Chem. Soc. 1988, 110 (12), 4063–4065.

(99)

(86)

Green, M. M.; Khatri, C. A.; Reidy, M. P.; Levon, K. DiluteSolution Chiral Optical Changes Signal the Thermally Reversible Gelation of Poly(n-Hexyl Isocyanate) in Hydrocarbon Solvents. Macromolecules 1993, 26 (17), 4723– 4725.

Brunsveld, L.; Vekemans, J. a J. M.; Hirschberg, J. H. K. K.; Sijbesma, R. P.; Meijer, E. W. Hierarchical Formation of Helical Supramolecular Polymers via Stacking of HydrogenBonded Pairs in Water. Proc. Natl. Acad. Sci. 2002, 99 (8), 4977–4982.

(100)

(87)

Green, M. M.; Peterson, N. C.; Sato, T.; Teramoto, A.; Cook, R.; Lifson, S. A Helical Polymer with a Cooperative Response to Chiral Information. Science 1995, 268 (5219), 1860–1866.

Kaiser, T. E.; Stepanenko, V.; Würthner, F. Fluorescent JAggregates of Core-Substituted Perylene Bisimides: Studies on Structure−Property Relationship, Nucleation−Elongation Mechanism, and Sergeants-andSoldiers Principle. J. Am. Chem. Soc. 2009, 131 (19), 6719– 6732.

(101)

Anderson, T. W.; Sanders, J. K. M.; Pantoş, G. D. The Sergeants-and-Soldiers Effect: Chiral Amplification in Naphthalenediimide Nanotubes. Org. Biomol. Chem. 2010, 8 (19), 4274.

(102)

Kulkarni, C.; Munirathinam, R.; George, S. J. Self-Assembly

(88)

Van Gestel, J.; Palmans, A. R. A.; Titulaer, B.; Vekemans, J. A. J. M.; Meijer, E. W. “Majority-Rules” operative in Chiral Columnar Stacks of C3-Symmetrical Molecules. J. Am.

13

ACS Paragon Plus Environment

Journal of the American Chemical Society 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Supramolecular Block Copolymers by Kinetically Controlled Co-Self-Assembly of Planar and Core-Twisted Perylene Bisimides. Nat. Commun. 2015, 6 (May), 1–8.

of Coronene Bisimides: Mechanistic Insight and Chiral Amplification. Chem. Eur. J. 2013, 19 (34), 11270–11278. (103)

Wang, F.; Gillissen, M. A. J.; Stals, P. J. M.; Palmans, A. R. A.; Meijer, E. W. Hydrogen Bonding Directed Supramolecular Polymerisation of Oligo(phenyleneEthynylene)s: Cooperative Mechanism, Core Symmetry Effect and Chiral Amplification. Chem. Eur. J. 2012, 18 (37), 11761–11770.

(104)

Helmich, F.; Smulders, M. M. J.; Lee, C. C.; Schenning, A. P. H. J.; Meijer, E. W. Effect of Stereogenic Centers on the Self-Sorting, Depolymerization, and Atropisomerization Kinetics of Porphyrin-Based Aggregates. J. Am. Chem. Soc. 2011, 133 (31), 12238–12246.

(105)

Smulders, M. M. J.; Stals, P. J. M.; Mes, T.; Paffen, T. F. E.; Schenning, A. P. H. J.; Palmans, A. R. A.; Meijer, E. W. Probing the Limits of the Majority-Rules Principle in a Dynamic Supramolecular Polymer. J. Am. Chem. Soc. 2010, 132 (2), 620–626.

(106)

Wilson, A. J.; van Gestel, J.; Sijbesma, R. P.; Meijer, E. W. Amplification of Chirality in Benzene Tricarboxamide Helical Supramolecular Polymers. Chem. Commun. 2006, 42, 4404–4406.

(107)

Roche, C.; Sun, H.-J.; Prendergast, M. E.; Leowanawat, P.; Partridge, B. E.; Heiney, P. A.; Araoka, F.; Graf, R.; Spiess, H. W.; Zeng, X.; et al. Homochiral Columns Constructed by Chiral Self-Sorting During Supramolecular Helical Organization of Hat-Shaped Molecules. J. Am. Chem. Soc. 2014, 136 (19), 7169–7185.

(108)

Wolffs, M.; Van Velthoven, J. L. J.; Lou, X.; Bovee, R. A. A.; Pouderoijen, M.; Van Dongen, J. L. J.; Schenning, A. P. H. J.; Meijer, E. W. Influence of the Solvent and the Enantiomeric Purity on the Transition between Different Supramolecular Polymers. Chem. Eur. J. 2012, 18 (47), 15057–15064.

(109)

Korevaar, P. A.; Grenier, C.; Markvoort, A. J.; Schenning, A. P. H. J.; de Greef, T. F. A.; Meijer, E. W. Model-Driven Optimization of Multicomponent Self-Assembly Processes. Proc. Natl. Acad. Sci. 2013, 110 (43), 17205–17210.

(110)

Buendía, J.; García, F.; Yélamos, B.; Sánchez, L. Transfer and Amplification of Chirality in Phe-Based: C3-Symmetric Non-Ionic Amphiphiles. Chem. Commun. 2016, 52 (57), 8830–8833.

(111)

Korevaar, P. A.; Grenier, C.; Meijer, E. W. Toward ModelDriven Engineering of Supramolecular Copolymers. J. Polym. Sci. Pt. A Polym. Chem. 2015, 53 (2), 385–391.

(112)

García-Iglesias, M.; de Waal, B. F. M.; de Feijter, I.; Palmans, A. R. A.; Meijer, E. W. Nanopatterned Superlattices in Self-Assembled C2-Symmetric Oligodimethylsiloxane-Based Benzene-1,3,5Tricarboxamides. Chem A Eur. J. 2015, 21 (1), 377–385.

(113)

Thota, B. N. S.; Lou, X.; Bochicchio, D.; Paffen, T. F. E.; Lafleur, R. P. M.; van Dongen, J. L. J.; Ehrmann, S.; Haag, R.; Pavan, G. M.; Palmans, A. R. A.; et al. Supramolecular Copolymerization as a Strategy to Control the Stability of Self-Assembled Nanofibers. Angew. Chem. Int. Ed. 2018, 57 (23), 6843–6847.

(114)

Petkau-Milroy, K.; Sonntag, M. H.; Van Onzen, A. H. A. M.; Brunsveld, L. Supramolecular Polymers as Dynamic Multicomponent Cellular Uptake Carriers. J. Am. Chem. Soc. 2012, 134 (19), 8086–8089.

(115)

Müller, M. K.; Brunsveld, L. A Supramolecular Polymer as a Self-Assembling Polyvalent Scaffold. Angew. Chem. Int. Ed. 2009, 48 (16), 2921–2924.

(116)

Görl, D.; Zhang, X.; Stepanenko, V.; Würthner, F.

Page 14 of 21

(117)

Das, A.; Vantomme, G.; Markvoort, A. J.; ten Eikelder, H. M. M.; Garcia-Iglesias, M.; Palmans, A. R. A.; Meijer, E. W. Supramolecular Copolymers: Structure and Composition Revealed by Theoretical Modeling. J. Am. Chem. Soc. 2017, 139 (20), 7036–7044.

(118)

Boott, C. E.; Leitao, E. M.; Hayward, D. W.; Laine, R. F.; Mahou, P.; Guerin, G.; Winnik, M. A.; Richardson, R. M.; Kaminski, C. F.; Whittell, G. R.; et al. Probing the Growth Kinetics for the Formation of Uniform 1D Block Copolymer Nanoparticles by Living Crystallization-Driven SelfAssembly. ACS Nano 2018, 12 (9), 8920–8933.

(119)

Baker, M. B.; Gosens, R. P. J.; Albertazzi, L.; Matsumoto, N. M.; Palmans, A. R. A.; Meijer, E. W. Exposing Differences in Monomer Exchange Rates of Multicomponent Supramolecular Polymers in Water. ChemBioChem 2016, 17 (3), 207–213.

(120)

Bakker, M. H.; Lee, C. C.; Meijer, E. W.; Dankers, P. Y. W.; Albertazzi, L. Multicomponent Supramolecular Polymers as a Modular Platform for Intracellular Delivery. ACS Nano 2016, 10 (2), 1845–1852.

(121)

Adelizzi, B.; Aloi, A.; Van Zee, N. J.; Palmans, A. R. A.; Meijer, E. W.; Voets, I. K. Painting Supramolecular Polymers in Organic Solvents by Super-Resolution Microscopy. ACS Nano 2018, 12 (5), 4431–4439.

(122)

Boott, C. E.; Laine, R. F.; Mahou, P.; Finnegan, J. R.; Leitao, E. M.; Webb, S. E. D.; Kaminski, C. F.; Manners, I. In Situ Visualization of Block Copolymer Self-Assembly in Organic Media by Super-Resolution Fluorescence Microscopy. Chem. Eur. J. 2015, 21 (51), 18539–18542.

(123)

Lou, X.; Lafleur, R. P. M.; Leenders, C. M. A.; Schoenmakers, S. M. C.; Matsumoto, N. M.; Baker, M. B.; van Dongen, J. L. J.; Palmans, A. R. A.; Meijer, E. W. Dynamic Diversity of Synthetic Supramolecular Polymers in Water as Revealed by Hydrogen/deuterium Exchange. Nat. Commun. 2017, 8, 15420.

(124)

Hendrikse, S. I. S.; Wijnands, S. P. W.; Lafleur, R. P. M.; Pouderoijen, M. J.; Janssen, H. M.; Dankers, P. Y. W.; Meijer, E. W. Controlling and Tuning the Dynamic Nature of Supramolecular Polymers in Aqueous Solutions. Chem. Commun. 2017, 53 (14), 2279–2282.

(125)

Wijnands, S. P. W.; Engelen, W.; Lafleur, R. P. M.; Meijer, E. W.; Merkx, M. Controlling Protein Activity by Dynamic Recruitment on a Supramolecular Polymer Platform. Nat. Commun. 2018, 9 (1).

(126)

Ellis, T. K.; Galerne, M.; Armao IV, J. J.; Osypenko, A.; Martel, D.; Maaloum, M.; Fuks, G.; Moulin, E.; Gavat, O.; Giuseppone, N. Supramolecular Electropolymerization. Angew. Chem. Int. Ed. 2018, 57, 15749-15753

(127)

Adhikari, B.; Suzuki, T.; Xu, L.; Yamauchi, M.; Karatsu, T.; Yagai, S. Photoresponsive Supramolecular Copolymers from Diarylethene–perylene Bisimide Hydrogen Bonded Complexes. Polymer 2017, 128, 356–362.

(128)

Hafner, R. J.; Tian, L.; Brauer, J. C.; Schmaltz, T.; Sienkiewicz, A.; Balog, S.; Flauraud, V.; Brugger, J.; Frauenrath, H. Unusually Long-Lived Photocharges in Helical Organic Semiconductor Nanostructures. ACS Nano 2018, 12 (9), 9116–9125.

14

ACS Paragon Plus Environment

Page 15 of 21 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Journal of the American Chemical Society

Table of Content

15

ACS Paragon Plus Environment

Journal of the American Chemical Society 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Figure 1 85x47mm (300 x 300 DPI)

ACS Paragon Plus Environment

Page 16 of 21

Page 17 of 21 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Journal of the American Chemical Society

Figure 2 170x76mm (300 x 300 DPI)

ACS Paragon Plus Environment

Journal of the American Chemical Society 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Figure 3 170x99mm (300 x 300 DPI)

ACS Paragon Plus Environment

Page 18 of 21

Page 19 of 21 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Journal of the American Chemical Society

Figure 4 85x131mm (300 x 300 DPI)

ACS Paragon Plus Environment

Journal of the American Chemical Society 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Figure 5 85x114mm (300 x 300 DPI)

ACS Paragon Plus Environment

Page 20 of 21

Page 21 of 21 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Journal of the American Chemical Society

Figure 6 170x153mm (300 x 300 DPI)

ACS Paragon Plus Environment