Читать книгу Supramolecular Polymers and Assemblies - Andreas Winter - Страница 4

List of Illustrations

Оглавление

1 Chapter 1Figure 1.1 Schematic representation of a polymer based on non‐covalent interactions.Figure 1.2 Schematic representation of the different types of supramolecular polymerization. Source: Winter et al. [39]. © 2012 Elsevier B.V.Figure 1.3 Representation of the theoretical DP as a function of the association constant (Ka in M−1) for a typical supramolecular polymerization according to an isodesmic model at two different concentrations. Source: Brunsveld et al. [27]. © 2001 American Chemical Society.Figure 1.4 Schematic representation of the three main mechanisms known for the supramolecular polymerization processes: (a) isodesmic, (b) ring‐chain mediated, and (c) cooperative supramolecular polymerization. Source: Winter et al. [39]. © 2012 Elsevier B.V. Figure 1.5 Schematic representation of the IDP in which the intermolecular equilibrium constant (K) is independent of the length of the assembly (the mechanism is shown for a bifunctional monomer of the Ia‐type, see also Figure 1.2). Source: Winter et al. [39]. © 2012 Elsevier B.V. Figure 1.6 (a) Schematic drawing of an energy diagram for an IDP (i: size of the oligomer, ΔG0: free energy in arbitrary units). (b) Evolution of the number‐ and weight‐averaged DP (<DP>N and <DP>W) and the dispersity (Đ) as a function of equilibrium constant and total concentration of monomer (K·ct). Source: de Greef et al. [26]. © 2009 American Chemical Society. Figure 1.7 Illustration of the characteristic properties of a temperature‐dependent IDP according to van der Schoot's model: (a) fraction of polymerized material (φ) vs. the dimensionless temperature T/Tm; (b) <DP>N vs. T/Tm. In both plots, the curves obtained for different enthalpies are shown (ΔHp = −30, −40, and −50 kJ mol−1, respectively). Source: van der Schoot et al. [57]. © 2005 Taylor & Francis. Figure 1.8 Illustration of the characteristic properties of a temperature‐dependent IDP according to the “free association” model: (a) fraction of polymerized monomers (φ) vs. T/Tm (assuming fully flexible polymer chains and a cubic lattice); (b) heat capacity at constant volume (CV) vs. T/Tm. In both plots, the curves obtained for various enthalpy (ΔHp = −30, −40, and −50 kJ mol−1, respectively) and entropy values (ΔSp = −100, −133, and −166 J mol−1 K−1, respectively) are shown; in all cases, the initial volume fraction of the monomers has been set to 0.1. Source: Modified from Dudowicz et al. [50]; Douglas et al. [51].Figure 1.9 Schematic representation of the generalized mechanism of a ring‐chain‐mediated supramolecular polymerization. The intermolecular binding constants (Kinter) are related to the intermolecular association of molecules, whereas the intramolecular binding constant Kintra(n‐mer) is assigned to the ring closure of monomers, oligomers, and polymers. Source: Winter et al. [39]. © 2012 Elsevier B.V. Figure 1.10 (a) Schematic representation of Kuhn's concept of effective concentration (ceff) for a heteroditopic oligomer (i.e. having two different end groups, A and B) [74]. In solution, the end group A will experience an effective concentration of B, if the latter one cannot escape from the sphere of radius l, which is identical to the length of the stretched chain. Thus, the intramolecular association between the termini becomes favored for ceff values higher than the actual concentration of B end groups. (b) Illustration of how the equilibrium concentration of chains and macrocycles can be correlated to the total concentration (ct) of a ditopic monomer in dilute solution; such a ring‐chain supramolecular polymerization typically features a critical concentration. Source: de Greef et al. [26]. © 2009 American Chemical Society.Figure 1.11 (a) Illustration of the fraction of polymerized monomer as a function of Kinter·ct for three different EM1 values and a fixed value of Kinter (106 M−1). (b) Illustration of the evolution of <DP>N as a function of Kinter·ct for various EM1 values. Source: Flory and Suter [91].Figure 1.12 Schematic representation of the formation of a poly(pseudorotaxane) via a ring‐chain equilibrium. Source: Cantrill et al. [95]. © 2001 American Chemical Society.Figure 1.13 Schematic representation of a typical cooperative supramolecular polymerization reaction (nucleation‐elongation mechanism). Kn and Ke represent the association constants for the nucleation and the elongation phase, respectively (Kn < Ke). Source: Winter et al. [39]. © 2012 Elsevier B.V.Figure 1.14 Schematic illustration of the energy diagrams of a cooperative nucleated (a) and a cooperative downhill supramolecular polymerization (b). In both plots, the axis of abscissae represents the oligomer's size (i), whereas the ordinate measures the ΔG0 in arbitrary units. In diagram (a), the size of the nucleus is 2 (i.e. dimeric nucleus); in diagram (b), a tetrameric nucleus is depicted. Source: de Greef et al. [26]. © 2009 American Chemical Society.Figure 1.15 Illustration of the various thermodynamic states in supramolecular polymerizations on Gibbs free energy landscape. Source: Sorrenti et al. [40]. Licenced under CC BY 3.0.

2 Chapter 2Figure 2.1 Schematic representation of the basic guanidinium–carboxylate interaction as well as of several guanidinium receptor. Source: Dietrich et al. [11]; Linton and Hamilton [12].Figure 2.2 Schematic representation of the binding of N‐acetyl‐protected α‐amino carboxylates to the receptor 6. Source: Schmuck [14].Figure 2.3 (a) Schematic representation of the dimerization of the self‐complementary guanidinium derivative 7[16]. (b) Schematic representation of the supramolecular self‐assembly of the heteroditopic derivatives 8. Source: Schmuck et al. [17]. Figure reproduced with kind permission. © 1999 Wiley‐VCH and 2000 American Chemical Society, respectively.Figure 2.4 (a) Schematic representation of the self‐assembly of two complementary components in a Vernier‐type fashion (the most straightforward case, i.e. combining a ditopic and a tritopic building block is depicted). (b) Schematic representation of the self‐assembly of 9 and 10 into a molecular [2×3]‐Vernier motif. Source: Kelly et al. [21].Figure 2.5 Schematic representation of the template‐driven self‐assembly of 11 and 12 into a supramolecular rectangle. Source: Terfort and von Kiedrowski [22].Figure 2.6 Schematic representation of star‐shaped assembly 13 and the corresponding X‐ray single crystal structure (R = CF3). Source: Kraft and Fröhlich [23]. © 1998 Royal Chemical Society.Figure 2.7 Schematic representation of the regioselective intramolecular photolysis reaction in the ion‐paired derivative 14. Source: Breslow et al. [24].Figure 2.8 (a) Schematic representation of the bowl‐shaped triple‐ions 15–18; (b) schematic representation (left) and space‐filling model of the ion pair 16 × 18 (right). Source: Grawe et al. [27]. Figure reproduced with kind permission. © 2002 American Chemical Society.Figure 2.9 (a) Schematic representation of metalloporphyrin 19 and calix[4]arene 20, as building blocks for supramolecular capsule formation. (b) Illustration of the simulated structure of the capsule (CHARMn 24.0). Source: Rehm and Schmuck [7]. © 2010 Royal society of chemistry.Figure 2.10 Schematic representation of cavitand 21 and its anion‐supported self‐assembly into a (212X4) capsule (X denotes as monovalent anion). Source: Oshovsky et al. [31]. © 2006 American Chemical Society.Figure 2.11 Proposed phase diagram for the supramolecular polymer formed by HOOC–PαMS–COOH (MW = 10 kDa) and H2N–PI–NH2PIP (MW = 18 kDa). The constituent blocks phase separated at the UCST. ODT denotes the regime for the order–disorder transition of the block copolymer structure. Tg and Ti define the glass transition and dissociation temperature, respectively. Microphase separation can be observed in the left‐to‐right diagonally hatched area; the mixture of telechelic polymers is macroscopically phase separated in the right‐to‐left diagonally hatched area. In the stippled regime, the copolymer is disordered phase, and finally, the clear area represents a homogeneous mixture of the two constituent polymers. The dashed lines represent the proposed continuations of the curves, which were experimentally not accessible due to ionic aggregation and/or cleavage of the supramolecular bonds. Source: Russell et al. [47]. © 1988 American Chemical Society.Figure 2.12 (a) Scheme representation of the vesicles formed by the self‐assembly of PS‐COOH and PNIPAM‐NH2 in aq. dioxane. (b) Representative TEM image of the thusly obtained vesicles. Source: Qian and Wu [64]. Figure reproduced with kind permission. © 2008 American Chemical Society.Figure 2.13 Transmission electron microscopy (TEM) images of ionically end‐capped PS‐b‐PI: (a) Me3N+‐PS‐b‐PI; (b) Me3N+ ‐ PS ‐ b ‐ PI ‐ SO3. Source: Schädler et al. [67]. Figure reproduced with kind permission. © American Chemical Society.Figure 2.14 Pictures of an “intelligent” supramolecular rubber, which exhibited (a) self‐healing and (b) shape‐memory properties. Source: Wang et al. [68]. Figure reproduced with kind permission. © 2015 The Royal Chemical Society.Figure 2.15 Schematic representation of the self‐assembly of a multi‐arm, star‐shaped polymers containing a POM, as polyanionic core (counterions omitted for clarity). Representative (a) TEM, (b) AFM, and (c) Scanning electron microscopy (SEM) images of the vesicles formed by the star‐shaped polymers are also shown (n = 169). Source: Zhang et al. [69]. Figure reproduced with kind permission. © 2012 The Royal Chemical Society.Figure 2.16 Schematic representation of POM‐centered supramolecular polymers via a surface‐started RAFT polymerization. As shown by representative TEM images, the morphology of the self‐assembled nanostructure depended on the length of the PS chains. Source: Cao et al. [72]. Figure reproduced with kind permission. © 2016 The Royal Chemical Society.Figure 2.17 Schematic representation of the morphologies of pristine PS488b‐P4VP95 (a) and the nanocomposite with added H4SiW12O40 (b). Source: Zhang et al. [76]. Figure reproduced with kind permission. © 2016 Elsevier B.V. Figure 2.18 Schematic representation of the self‐assembly of the heteroditopic monomer 22 in a head‐to‐tail fashion. The concentration‐dependency of the self‐assembly process, as studied by 1H NMR spectroscopy, is also shown. Source: Schmuck [20]. © 2001 Elsevier B.V.Figure 2.19 Schematic representation of the different self‐assembly modes of zwitterion 23 dictated by the length of the alkyl chain. Source: Schmuck et al. [79]. © 2007 American Chemical Society.Figure 2.20 (a) Schematic representation of zwitterions 24 functionalized with amino‐acid residues. (b) Representative TEM image of a vesicle formed by the self‐assembly of 24a in DMSO (after staining with uranyl acetate). (c) Calculated structure of a membrane segment (the Me‐groups of the alanine moieties, essential for the vesicle formation are depicted in yellow); a view along the rows of stacked dimers showing the vesicle curvature as well as side view of the stacked dimers showing their alternating antiparallel orientation. Source: Rehm et al. [80]. Figure reproduced with kind permission. © 2008 American Chemical Society.Figure 2.21 Schematic representation of the tris‐zwitterion 25. The AFM images of the assemblies after spin‐coating onto mica substrates are also depicted: large plates were obtained on the surface at high rotational speed (7000 rpm, a and b); the plates merged at lower rotational speed (5000 rpm, c) and, finally, fully disintegrate into the 2D network of ribbons (d). A cross‐sectional plot of three representative plates is also shown (e, also depicted as yellow line in b) exhibited uniform heights and diameters of c. 450 and 2.5 nm, respectively. Source: Rehm et al. [82]. Figure reproduced with kind permission. © 2012 The Royal Chemical Society. Figure 2.22 Polarizing microscopy image of the PAA–dodecyltrimethylammonium complex (extension of λ = 1.5 nm). Source: Antonietti and Conrad [87]. Figure reproduced with kind permission. © 1994 Wiley‐VCH.Figure 2.23 Schematic representation of the supramolecular grafting of sulfonates onto P4VP. Source: Ikkala et al. [93].Figure 2.24 Schematic representation of the formation of helical strands from the achiral polyacetylene derivative 25 induced by ion pairing with chiral amines. Source: Yashima et al. [101]. © 1999 Springer Nature.Figure 2.25 (a) Schematic representation of polymers R28 and 29 that self‐assemble into right‐handed double helices in polar‐aprotic solvents. (b) Wide‐angle XRD pattern and AFM phase image (40 nm × 40 nm) of R28 × 29 along with a molecular model of the double‐stranded helical structure. Source: Maeda et al. [103]. Figure reproduced with kind permission. © 2008 American Chemical Society. Figure 2.26 Schematic representation of the two different structural modes for stoichiometric PECs. (a) Ladder‐type structure and (b) “Scrambled egg” structure.

3 Chapter 3Figure 3.1 Schematic representation of the various types of architectures accessible via H‐bonding interactions (A: H‐bonding acceptor, D: H‐bonding donor) [3]. Source: Redrawn from Binder and Zirbs [3]. © 2007 Springer Nature. Figure 3.2 (a) Classification of H‐bonds according to Jeffrey and Saenger (D: H‐bonding donor, A: H‐bonding acceptor, M: metal ion). Source: Jeffrey and Saenger [16]. © 1991 Springer Nature.Figure 3.3 Schematic representation of various single H‐bonding motifs.Figure 3.4 Schematic representation of various two‐centered H‐bonding motifs (A: adenine, T: thymine, G: guanine, C: cytosine).Figure 3.5 Schematic representation of various triple H‐bonding motifs (A: adenine, T: thymine, G: guanine, C: cytosine).Figure 3.6 Schematic representation of triple H‐bonding arrays exhibiting different Ka values (D: H‐bonding donor; A: H‐bonding acceptor, KA: association constant in CHCl3). Source: Brunsveld et al. [5].Figure 3.7 (a) Schematic representation of various quadruple H‐bonding motifs (from left to right). Source: Refs. [18,19] and Prabhakaran et al. [20]Figure 3.8 Schematic representation of Napy‐induced translation of homodimeric into heterodimeric assemblies via quadruple H‐bonding. Two examples according to Corbin and Zimmerman (a) and Chen (b) are shown in [21,37].Figure 3.9 Schematic representation of heterodimers based on sextuple H‐bonding systems. Source: (a) Chang and Hamilton [38] and (b) Yang et al. [41].Figure 3.10 The utilization of single H‐bonding for the formation of main‐chain supramolecular materials (left) and inter‐chain connection polymer blends (right). Source: Redrawn from Binder and Zirbs [3]. © 2007 Springer Nature.Figure 3.11 The formation of supramolecular ladder‐type polymers and networks based on monomers 1 and 2. Source: St.Pourcain and Griffin [73].Figure 3.12 Schematic representation of the formation of a supramolecular polymeric network using double H‐bonding interactions of the self‐complementary uradiazole units.Figure 3.13 (a) Schematic representation of the telechelic PDMSs 3 as monomers for the supramolecular ring‐chain equilibrium polymerization. (b) FT‐IR spectra of (3b)n at two different concentrations (c = 1.4 and 20 g l−1). (c) Concentration dependence of the reduced specific viscosity of (3b)n and the corresponding dibenzyl ester Bn–3b–Bn (Bn = benzyl, hexane, 25 °C). Source: Abed et al. [4]. Figure reproduced with kind permission. © 2000 American Chemical Society.Figure 3.14 (a) Schematic representation of the chiral two‐centered H‐bonding unit 4 with its bicyclo[3.3.1]nonane core. (b) Schematic representation of the structure‐dependent supramolecular self‐assembly of monomers 5 and 6.Figure 3.15 (a) Schematic representation of the supramolecular polymer (7)n. (b) Schematic representation of the dense 1D packing of 7a in a slipped fashion (the red twisted blocks represent the PBI core, the bay substituents are shown in gray cones with a blue apex and H‐bonding interactions are indicated as green lines). Source: Würthner et al. [101]. Figure reproduced with kind permission. © 2016 American Chemical Society. Figure 3.16 Schematic representation of the formation of parent DPP from its Nt‐Boc‐protected derivative 8; DPP self‐assembles into linear supramolecular polymer due to double H‐bonding interactions. Figure 3.17 (a) Schematic representation of the Nt‐Boc‐protected DPP oligomers 9 and 10. (b) Schematic representation of the Nt‐Boc‐protected DPP and PBI dyes 11 and 12. Figure 3.18 Schematic representation of the supramolecular polymerization of the homotelechelic monomers 13a and 13b into an alternating copolymer. Source: Fouquey et al. [2].Figure 3.19 Schematic representation of the LC supramolecular polymer14. Source: Kotera et al. [115].Figure 3.20 (a) Schematic representation of the formation of supramolecular gels based on triple H‐bonding motifs in concert with additional non‐covalent interchain interactions. Representative field‐emission scanning electron microscopy (FE‐SEM) images of the fibers obtained from the dried benzene gels are also shown (left: 15a, right: 15b). (b) Schematic representation of monomer 16 used for the supramolecular polymerization with N‐dodecyl cyanuric acid. Source: Yagai et al. [119]. Figure reproduced with kind permission; © 2005 American Chemical Society. Figure 3.21 Schematic representation of the equilibrium between supramolecular structures of linear tape‐like and cyclic rosette‐like shapes.Figure 3.22 Schematic representation of the supramolecular polymer 17 based on triple H‐bonding interactions, supported by π-π stacking hydrophobic interactions and a PBI core. A representative TEM image of the mesoscopic superstructures that were fabricated by evaporation of a methylcyclohexane solutions of 17 (c = 5 × 10−5 mol l−1) on a carbon‐Formvar‐coated nickel grid (200 mesh) and subsequent staining with uranyl acetate is also shown. Source: Würthner et al. [130]. Figure reproduced with kind permission; © 1999 Wiley‐VCH.Figure 3.23 (a) Schematic representation of the heteroditopic PBI 18. (b) Schematic representation of the supramolecular polymerization of 18 into cylindrical nanostructures. (c) Representative TEM image of the nanocylinders. Source: Sinks et al. [133]. Figure reproduced with kind permission. © 2005 American Chemical Society. Figure 3.24 (a) Scanning tunnelling microscopy (STM) image of the honeycomb structure on a Ag‐layered Si substrate (inset: high‐resolution view of the substrate surface) [134]. (b) STM image of the Lu@C82‐loaded network on the Au(111) surface (area of 55 × 40 nm2). Source: Silly et al. [135]. Figure reproduced with kind permission; © 2003 Nature Publishing Group and 2008 The Royal Chemical Society.Figure 3.25 Schematic representation of the stepwise fabrication of a hybrid structure starting from a supramolecular honeycomb PBI–melamine network. Source: Silien et al. [139]. Figure reproduced with kind permission; © 2009 Wiley‐VCH. Figure 3.26 Schematic representation of the supramolecular self‐assembly of 20a and 21 into ribbon‐like aggregates. Representatively, a field‐emission SEM image of the aggregates is also shown. Source: Yagai et al. [123]. Figure reproduced with kind permission. © 2007 American Chemical Society. Figure 3.27 Schematic representation of the monomers 22, which self‐assembled into toroids and linear fibrils (a) or, upon mixing, generated a supramolecular helical polymer (b). Source: Aratsu et al. [141]. Figure reproduced with kind permission. © 2020 Springer Nature.Figure 3.28 A chain‐extended supramolecular block‐copolymer based on triple H‐bonding arrays. PIB: poly(iso‐butylene), PEK: poly(etherketone). Source: Binder et al. [146]. Figure reproduced with kind permission. © 2005 Wiley‐VCH.Figure 3.29 Schematic representation of the homotelechelic PPOs23. The pictures show pure 23a, pure 23b, and an equimolar mixture of both (from left to right). Source: Cortese et al. [148]. Figure reproduced with kind permission. © 2012 American Chemical Society.Figure 3.30 Schematic representation of the reversible self‐assembly of the nucleobase‐functionalized oligo(phenylene‐ethynylene)s 24 (A: adenine, T: thymine). The optical micrograph of an annealed film of the polymer (24a···24b···)n at 130 °C is also shown (100× magnification, the inset shows a section with a 500× magnification). Source: Figure reproduced with kind permission; © 2003 The Royal Society of Chemistry. Figure 3.31 (a) Schematic representation of the bolaamphiphiles 25 (for the structures of adenine [A] and thymine [T], see Figure 3.30). (b–d) Energy‐filtered transmission electron microscopy (EF‐TEM) images of the fibers formed by self‐assembly in 10% ethanolic aqueous solution at 25 °C (b: 25a, c: 25c, d: 1 : 1 ratio of 25b and 25c; the scale bar represents always 1 μm). Source: Modified from Shimizu et al. [151]. Figure reproduced with kind permission. © 2001 American Chemical Society.Figure 3.32 A highly flexible material was formed via the self‐assembly of α,ω‐dinucleobase‐functionalized PTHF (e.g. 26a with two N4‐(4‐tert‐butylbenzoyl)‐cytosine moieties). Source: Modified from Sivakova et al. [152]. Figure reproduced with kind permission. © 2005 American Chemical Society.Figure 3.33 Schematic representation of two important self‐complementary quadruple H‐bonding arrays with high Ka values. Figure 3.34 (a) Schematic representation of the UPy‐functionalized homotelechelic monomers27. (b) Representation of the critical concentrations of 20 in supramolecular ring‐chain equilibrium polymerizations. Source: ten Cate and Sijbesma [11]. Figure reproduced with kind permission. © 2003 Wiley‐VCH.Figure 3.35 Schematic representation of the UPy‐functionalized homotelechelic monomers 2931.Figure 3.36 Images of a bis‐UPy‐functionalized poly(ɛ‐caprolactam) that was processed into different scaffolds: (a) thin films, (b) fibers, (c) meshes (a representative SEM image is depicted), and (d) grids. Source: Dankers et al. [165]. Figure reproduced with kind permission. © 2005 Nature Publishing Group. Figure 3.37 Schematic representation of the fabrication of vascular graft from UPy‐equipped macromolecular components. The pictures show the results of the in‐vivo cell‐adhesion experiments. Source: van Alme et al. [166]. Figure reproduced with kind permission. © 2016 Wiley‐VCH.Figure 3.38(a) Schematic representation of the H‐bonded dimer 32. (b) SEM image of the surface roughness (scale bar: 2 μm). (c) Representation of the superhydrophobicity of the resulting surface: Water contact angle of c. 150°. Source: Han et al. [175]. Figure reproduced with kind permission. © 2004 American Chemical Society.Figure 3.39 Schematic representation of the formation of supramolecular AB‐type diblock copolymers due to complementary H‐bonding interactions. Source: Feldman et al. [178]. © 2008 American Chemical Society.Figure 3.40 Schematic representation of the UPy‐ and Napy‐functionalized homopolymers33 and 34. The microscopy images show the microstructure of the blends after annealing at various temperatures. Source: Modified from [178]. Figure reproduced with kind permission. © 2008 American Chemical Society.Figure 3.41 Schematic representation of the supramolecular diblock copolymers35 and 36. A representative TEM image of 35 showing the microphase separation in the solid state is also shown. Source: Rao et al. [186]. Figure reproduced with kind permission. © 2012 The Royal Chemical Society.Figure 3.42 Schematic representation of the formation of a strictly alternating supramolecular copolymer using the non‐self‐complementary monomers37 and 38. Source: Based on Park and Zimmerman [187]. © 2006 American Chemical Society.Figure 3.43 Schematic representation of the heteroditopic H‐bonding units 3941.Figure 3.44 Schematic representation of the photoreversible formation of the supramolecular polymer (c42)n.Figure 3.45 (a) Schematic representation of the various bis‐UPy‐functionalized π‐conjugated chromophores4345. (b) Schematic representation of the proposed energy‐transfer process within the supramolecular copolymers. (c) Representation of the emission colors of the three individual homopolymers and the three‐component copolymer in solution (top) as well as in the solid state (bottom). Source: Abbel et al. [195]. Figure reproduced with kind permission. © 2009 American Chemical Society.Figure 3.46 Schematic representation of the quadruple H‐bonding arrays 46 and 47 based on self‐complementary 1,3,5‐triamino‐2,4,6‐triazine derivatives (R denotes aliphatic and oligo(ethylene glycol) chains – chiral and non‐chiral ones).Figure 3.47 (a, b) Schematic representation of the self‐assembly of the heterodifunctional H‐bonding array48 into a supramolecular polymer. (c) Temperature‐dependent UV/vis absorption spectra of 48 (0.10 mM in decaline); the inset shows the evolution of the absorbance at 349 nm when the temperature was changed. Source: Ikeda et al. [197]. Figure reproduced with kind permission. © 2005 Wiley‐VCH.Figure 3.48 (a) Schematic representation of the linear supramolecular polymer49. (b) Representative electron‐microscopy images of the extended, densely packed fiber networks obtained from the solution of 49 (2.5 mmol l−1 in CHCl3/heptane, 1 : 4 ratio). Source: Berl et al. [198]. Figure reproduced with kind permission. © 2002 Wiley‐VCH.Figure 3.49 Schematic representation of the formation of homopolymers and diblock copolymers equipped with Hamiltonian receptors or barbiturate units. PS: polystyrene, PMMA: poly(methyl methacrylate), PLMA: poly(lauryl methacrylate), PnBuA: poly(n‐butyl acrylate). These macromonomers were assembled into supramolecular pseudo‐block copolymers via sextuple H‐bonding interactions. Source: Chen et al. [201]. Figure reproduced with kind permission. © 2012 The Royal Chemical Society. Figure 3.50 Schematic representation of the formation of a supramolecular helix–helix diblock copolymer. Source: Croom et al. [204]. Figure reproduced with kind permission. © 2016 American Chemical Society.Figure 3.51 Schematic representation of the heterotelechelic polymer50 that showed a concentration‐dependent self‐assembly behavior (i.e. formation of single‐chain macrocycles vs. chain‐extended polymers). Source: Altintas et al. [206]. Figure reproduced with kind permission. © 2010 The Royal Chemical Society.Figure 3.52 Schematic representation of Lehn's concept of double dynamic supramolecular polymers (“double dynamers). Source: Kolomiets et al. [209]. © 2005 The Royal Chemical Society.Figure 3.53 Schematic representation of the directed self‐assembly of the calix[4]arenes 51 and 52 into columnar structures via tris‐aminotriazine/barbiturate interactions. Source: Klok et al. [211]. Figure reproduced with kind permission. © 1999 American Chemical Society.Figure 3.54 Schematic representation of the supramolecular diblock copolymer53 featuring a sextuple H‐bonding linkage (PS block: Mn = 20 000 g mol−1, PEG block: Mn = 5000 g mol−1). A representative AFM image of a spin‐coated thin film is also depicted. Source: Yang et al. [41]. Figure reproduced with kind permission. © 2004 Wiley‐VCH.Figure 3.55 (a) Schematic representation of the cyclic eight‐residue peptide54 and of its self‐assembly into supramolecular nanotubes. (b) b1: TEM image of the supramolecular nanotubes adsorbed onto a carbon support film; b2: Low‐dose cryo‐microscopy image of a single nanotube. (c) Schematic representation of the cyclic four‐residue peptide 55. Source: Khazanovich et al. [217]. Figure reproduced with kind permission. © 1994 American Chemical Society.Figure 3.56 (a) Schematic representation of the formation of a “supramolecular rubber,” based on multiple H‐bonding. (b) Representation of a polymer network formed by mixtures of ditopic (blue) and tritopic (red) building blocks, self‐assembled by directional interactions (dotted lines). (c) Stress–strain curve of the supramolecular rubber. Source: Cordier et al. [230]. Figure reproduced with kind permission. © 2008 Springer Nature.

4 Chapter 4Figure 4.1 Schematic representation of the general architectures of metal‐containing polymers. Source: Wild et al. [67]. © 2011 The Royal Chemical Society. Figure 4.2 Schematic representation of the three main types of metal‐to‐ligand interactions. (a) Coordinative bonding; (b) ionic bonding; and (c) arene π‐complexation. Source: Refs. [2,46,84].Figure 4.3 Schematic representation of the different general methods utilized for the synthesis of metal‐containing (co)polymers. Source: Refs. [2,61].Figure 4.4 (a) Schematic representation of metal complexes containing pyridine‐based ligands of increasing denticity. (b) The Irving–Williams series.Figure 4.5 Schematic representation of the coordination polymers 1–3 showing an increased stability in solution due to enhanced metal‐to‐ligand interactions.Figure 4.6 Schematic representation of the synthesis of the heterobimetallic coordination polymer 4.Figure 4.7 Illustration of the sonication‐induced decrease in the molar mass of 2 (1.5 mM in toluene): (a) Molar mass distribution of samples taken during sonication. (b) Evolution of the MW value over five cycles of sonication (1 hour) followed by equilibration (23 hours). Source: Paulusse and Sijbesma [94]. Figure reproduced with kind permission. © 2004 Wiley‐VCH. Figure 4.8 (a) Schematic representation of the synthesis of coordination polymer 5 via the coordination of 1,4‐benzenediisocyanide to the [Pt33‐CO)(μ‐dppm)3]2+ cluster (the counterions are omitted for clarity, L^L: dppm). A representation of the X‐ray single‐crystal structure of the precursor cluster is also shown (here, only the phenyl C‐atoms bonded to each P‐atom are shown for clarity). Source: Bradford et al. [111]. Figure reproduced with kind permission. © 1994 American chemical society.Figure 4.9 Schematic representation of the synthesis of the metallopolymer 7 with linear Pt4 clusters within the main chain.Figure 4.10 Schematic representation of the synthesis of coordination polymer 8 featuring A‐frame‐units within the backbone.Figure 4.11 Schematic representation of the synthesis of 9, as a polymer‐analogs to cisplatin.Figure 4.12 Schematic representation coordination polymers through the center of the porphyrin rings. (a) Shish‐Kebab‐type assembly, (b and c) assemblies based on AB‐type monomers.Figure 4.13 Schematic representation of Shish–Kebab‐type metallo‐supramolecular polymers 10 and 11 incorporating phthalocyanine.Figure 4.14 Schematic representation of the synthesis of the metalloporphyrin‐containing polymer12 via a Glaser‐type polyaddition (a) and of the metallo‐supramolecular 2D ladder‐shaped assembly featuring enhanced charge mobility (b). Source: Taylor and Anderson [123]. © 1999 American chemical society.Figure 4.15 Schematic representation of the synthesis of a zigzag‐shaped metallo‐supramolecular coordination polymer. Source: You and Würthner [129]. © 2004 American Chemical Society. Figure 4.16 Schematic representation of the self‐assembly of a 4‐pyridyl‐substituted porphyrin derivative with Zn(II) ions toward 14.Figure 4.17 Schematic representation of the linear Au(I)‐containing metallopolymers 15, which exhibited an odd–even effect with respect to the chain conformation.Figure 4.18 Schematic representation of various metallopolymers containing ditopic NHC‐type ligands. (a) Monodentate and (b) bidentate binding.Figure 4.19 Schematic representation of the synthesis of the rigid metallopolymers 19.Figure 4.20 Schematic representation of the synthesis of the flexible metallo‐supramolecular polymer 20.Figure 4.21 Schematic representation of the synthesis of metallopolymers 21 and 23 based on tetrahedral bis‐bidentate complexes (a) and a phosphine‐based ligand (b), respectively.Figure 4.22 Schematic representation of the chemical structure of the metallo‐supramolecular polymer 22, an illustration of the proposed helical 3D‐conformation is also shown. Source: Kaes et al. [175]. © 1998 Wiley‐VCH.Figure 4.23 Schematic representation of the metallopolymers 24 and 25. The picture shows the green color of 24a that could be changed reversibly to colorless upon electrochemical reduction of the Cu(II) centers. Source: Hossain et al. [181]. Figure reproduced with kind permission. © 2013 Wiley‐VCH.Figure 4.24 Schematic representation of the synthesis of coordination polymers 26 and 27 containing bridging [Ni(mnt)2]2− units.Figure 4.25 Generalized schematic representation of the self‐assembly of ditopic bis‐tridentate ligands and divalent transition metal ions into linear metallo‐supramolecular polymers. Source: Redrawn from Chiper et al. [52]. 2009 John Wiley & Sons.Figure 4.26 Schematic representation of the synthesis of metallopolymers bearing amido‐ or imido‐linkages via two different routes (denoted as methods III and Vb, respectively, see also Figure 4.3).Figure 4.27 Schematic representation of the synthesis of metallo‐supramolecular assemblies from a flexible bis‐terpyridine ligand and Ru(II) ions.Figure 4.28 Schematic representation of the Ru(II)‐containing chain‐extended polymer 31.Figure 4.29 Schematic representation of the general structure of a metallo‐supramolecular homo as well as a diblock copolymer. Source: Redrawn from Schubert et al. [186].Figure 4.30 Schematic representation of the synthesis of polymers equipped with designated metal‐binding sites via the end‐functionalization and the initiator route. Source: Hoogenboom and Schubert [233]. © 2006 The Royal Chemical Society. Figure 4.31 LCST behavior of PNIPAM, tpy‐PNIPAM (32), and [Fe(32)2]X2 with different counterions X (i.e. Cl, AcO, and PF6) in water as monitored by transmittance measurements. Images of the vials below (top) and above the LCST (bottom) are also shown for 32 (left) and [Fe(32)2]2+ (right). Source: Chiper et al. [239]. © 2008 Wiley‐VCH. Figure 4.32 (a) Transmission electron microscopy (TEM) image of PEG–[Ru]–PS micelles in water (without staining). Source: Lohmeijer et al. [216]. Figure reproduced with kind permission. © 2004 The Royal Chemical SocietyFigure 4.33 Representative AFM images of a spin‐coated film (75 nm thickness) of PS375–[Ru]–PEG225 on a Si substrate, before (a) and after (b) the removal of the hydrophilic block. (c) XPS signature (C1s regime) of the film before (solid line) and after treatment with the oxidizing agent. Source: Fustin et al. [220]. Figure reproduced with kind permission. © 2005 Wiley‐VCH.Figure 4.34 Schematic representation of the macroligands33 and 34, which gave an A–[M]–B diblock copolymer in the presence of Zn(II) ions. The diblock copolymer assembled further into spheres and fibers as a function of m/n ratio (m: length of the PEG block, n: length of the P3HT block). Typical TEM images for both cases (i.e. low and high contents of P3HT, respectively) are also shown. Source: He et al. [247]. Figure reproduced with kind permission. © 2017 American Chemical Society.Figure 4.35 Schematic representation of the synthesis of a metallo‐supramolecular polymer with pending MPEG and poly(2‐diisopropylaminoethyl methacrylate) (PDPA) chains on each repeat unit. This material self‐assembled into striped nanosheets in water. A representative TEM image of such a nanosheet is also depicted. Source: Zhang et al. [250]. Figure reproduced with kind permission. © 2020 The Royal Chemical Society. Figure 4.36 Schematic representation of the metallo‐supramolecular modification of tpy‐functionalized micelles via grafting‐onto and cross‐linking methodologies. Source: Refs. [224,251]. © 2008 and 2009 The Royal Chemical Society. Figure 4.37 Schematic representation of the synthesis of the metallo‐supramolecular ABA‐type triblock copolymers 35.Figure 4.38 Schematic representation of the metallo‐supramolecular polymerization of the ditopic ligand 36 in the presence of transition and rare‐earth metal ions. The pictures illustrate the thermoresponsive (a) and the thixotropic behavior (b) of representative combinations of metal ions. Source: Beck and Rowan [253]. Figure reproduced with kind permission. © 2003 American Chemical Society.Figure 4.39 Schematic representation of the metallo‐supramolecular polymerization of rigid bis‐tridentate ligands with Ru(II) ions.Figure 4.40 Schematic representation of the metallo‐supramolecular polymers reported by the Kurth and Higuchi groups. The colors of the metallopolymers in dilute solution are also shown [a: Fe(II), b: Ru(II), and c: Co(II)]. Source: Han et al. [318]. Figure reproduced with kind permission. © 2004 American Chemical Society.Figure 4.41 Representation of the hierarchical self‐assembly of a metallo‐supramolecular polymer by electrostatic self‐assembly processes. (a) Formation of layer‐by‐layer assemblies, (b) formation of core‐shell or hollow particles and (c) formation of metallopolymer‐amphiphile complexes. PSS: poly(styrene sulfonate); PEI: polyethyleneimine; DHP: dihexadecyl phosphonate [298,299,308]. Source: Wild et al. [67]. © 2011 The Royal Chemical Society.Figure 4.42 Schematic representation of selected metallo‐supramolecular polymers assembled from Zn(II) ions and π‐conjugated bis‐terpyridine ligands.Figure 4.43 (a) Mixing triangle of the metallo‐supramolecular polymers 38b, 38c, and 38d. (b) Picture of the corresponding solutions in a quartz microtiter plate (excitation at λ = 365 nm). (c) Position of the observed emission according to the CIE color scheme. Source: Wild et al. [287]. Figure reproduced with kind permission. © 2013 The Royal Chemical Society.Figure 4.44 Schematic representation of the chiral Ru(II)‐ and Fe(II)‐containing metallopolymers 39–42.Figure 4.45 Schematic representation of the bpp‐, btp‐, and pybox‐type ligands 43–46 used for the self‐assembly with Fe(II), Eu(III), Ru(II), and Zn(II) ions.Figure 4.46 Schematic representation of the “scorpionate”‐type metallopolymer [Fe(47)]n, the SEC trace (UV detector) of the purified polymer is also shown. Source: Qin et al. [366]. Figure reproduced with kind permission. © 2012 Wiley‐VCH.Figure 4.47 Schematic representation of the general structure of poly(metal acetylide)s and poly(metal arylide)s.Figure 4.48 Schematic representation of the polyplatinyne synthesis, according to Hagihara's original protocol.Figure 4.49 Schematic representation of selected examples of polyplatinynes highlighting the broad structural diversity reported in literature.Figure 4.50 Schematic representation of the poly(metal acetylide)s synthesis via metathesis reactions: (a) classical route and (b) extended one‐pot route.Figure 4.51 Schematic representation of Puddephatt's synthesis of Au(I)‐containing polymers 49 and 50.Figure 4.52 Schematic representation of the Au(I)‐containing monomer 51 which assembled into supramolecular fibers due to aurophilic and H‐bonding interactions. A representative TEM image of the obtained fiber network is also shown. Source: Chen et al. [405]. Figure reproduced with kind permission. © 2017 The Royal Chemical Society.Figure 4.53 Schematic representation of metallo‐supramolecular polymers based on pincer‐type binding motives according to (a) Loeb and Shimidzu [408] was well as (b) van Koten et al. [409].Figure 4.54 Schematic representation of two coordination polymers based on Pd(II)‐pincer complexes according to the Craig (51) and the Weck groups (52), respectively.Figure 4.55 (a) Schematic representation of the proposed synthesis of metallopolymer 56 featuring pincer‐type coordination and carbon‐to‐metal bonds. (b) Schematic representation of the metallopolymer 57 featuring bis(triazol‐4‐yl)pyridine coordination and carbon‐to‐metal bonds; a representation of the 3D SEC analysis is also depicted. Source: Schulze et al. [416]. Figure reproduced with kind permission. © 2017 American Chemical Society.Figure 4.56 (a) Schematic representation of metallo‐supramolecular polymers 58, based on Zn(II)‐dpm bis‐complexes. (b) Scanning electron microscopy (SEM) images of the assembly 58 containing an angular ligand in THF (top, inset: 25 μm × 25 μm fluorescence micrograph) and from a 2 : 1 THF/water mixture (bottom). Source: Maeda et al. [419]. Figure reproduced with kind permission; © 2006 American Chemical Society.Figure 4.57 Schematic representation of the Cu(II)‐containing chain‐extended polymers 59 showing LC behavior.Figure 4.58 Schematic representation of the synthesis of linear and branched coordination polymers, incorporating luminescent Znq2‐ or Alq3‐type complexes.Figure 4.59 Schematic representation of metallo‐supramolecular polymers 61 and 62.Figure 4.60 Schematic representation of metallo‐supramolecular polymers, based on salen (63) and catechol bis‐complexes (64).Figure 4.61 Structural characteristics of the coordination assemblies obtained from Ln(III) ions and 65, as a function of the concentration. Source: Vermonden et al. [443]. Figure reproduced with kind permission; © 2004 Wiley‐VCH.Figure 4.62 Schematic representation of the ditopic ligand, which was used for the complexation of Eu(III) ions. Photographs of the free‐standing Kn[Eu(66)]n polymer film upon excitation with 365‐nm‐light – under NEt3 vapor and upon exposure to HCl gas – are also shown. The comparison of the photoluminescence intensity reveals the switchability between the “ON” and “OFF” state. Source: Sato and Higuchi [445]. Figure reproduced with kind permission. © 2012 The Royal Chemical Society.Figure 4.63 Schematic representation of the synthesis of a white‐light‐emitting heterobimetallic supramolecular polymer. Source: Sato and Higuchi [448]. Figure reproduced with kind permission. © 2019 Elsevier B.V.Figure 4.64 (a) Schematic representation of the heterobimetallic coordination polymer 68. (b) Schematic representation of the tetrathiolate‐based metallopolymers 69 and their electric conductivities.Figure 4.65 Schematic representation of typical polymetallocene structures.Figure 4.66 Schematic representation of the synthesis of the ferrocene‐containing polymers70 (a) and 71 (b) by a metallo‐supramolecular polymerization. Source: Refs [465,466].Figure 4.67 Schematic representation of the generalized structures of polydecker sandwich complexes (Types A and B) as well as linear polymetallocenes (Types C and D). Source: Redrawn from Manners [41].Figure 4.68 Schematic representation of the synthesis of type‐A polydecker complexes incorporating (a) Ni(II) (72) and (b) Rh(II) centers (73). Figure 4.69 Schematic representation of the oligometallocene 74.Figure 4.70 Schematic representation of the synthesis of the homo‐ and heterometallic polydecker assemblies 75.Figure 4.71 Schematic representation of the type‐C polymetallocene 76.

5 Chapter 5Figure 5.1 Illustration of the interactions between two idealized p atoms as a function of their orientation – two “attractive” geometries and the “repulsive” face‐to‐face‐type geometry are shown. Source: Hunter and Sanders [4]; © 1990 American Chemical Society.Figure 5.2 Schematic representation of the hexa‐peri‐substituted hexabenzocoronenes1 along with the proposed stacking mode. The polarized optical microscopy (POM) (a) and atomic‐force microscopy (AFM) and (b) images of drop‐casted fibers of 1e are also shown. Source: Kastler et al. [12]. Figure reproduced with kind permission; © 2005 American Chemical Society.Figure 5.3 Schematic representation of the amphiphilic HBCs 2 and their supramolecular polymerization into tubular structures. Source: Modified from Hill et al. [13].Figure 5.4 Schematic representation of the self‐assembly of a discotic amphiphile in aqueous medium. Source: Aida [15], © 2020 John Wiley and Sons.Figure 5.5 Schematic representation of 3,5‐diphenylisoxazoles featuring a local dipole moment and a planar structure. Source: Haino et al. [17], © 2008 Royal Society of Chemistry.Figure 5.6 Schematic representation of the C3‐symmetric diphenylisoxazoles3 along with the proposed packing mode. FE‐SEM images of the fibers obtained from 3a (a) and 3b (b). Source: Haino et al. [17]. Figure reproduced with kind permission; © 2008 The Royal Chemical Society.Figure 5.7 (a) Schematic representation of the basic diketopyrrolopyrrole structures 4. (b) Schematic representation of the donor–acceptor–donor molecules 5, which assembled due to π–π stacking interactions of their outer electron‐rich chromophores. Source: Lee et al. [18], © 2011 John Wiley and Sons.Figure 5.8 Schematic representation of representative core‐unsubstituted PBI dyes (610).Figure 5.9 Schematic representation of representative bay‐substituted PBI dyes (11 and 12).Figure 5.10 (a) Representation of the color tunability of the emission of 6a in solution, the red emission from a spin‐coated film of 6a is shown on the very right. (b) POM image of the LC phase of 6a. Source: Chen et al. [25]. Figure reproduced with kind permission; © 2007 Wiley‐VCH.Figure 5.11 (a) Schematic representation of the proposed columnar stacking mode of PBI molecules. (b) Plot of the standard Gibbs free energy (ΔG0) for an isodesmic aggregation process vs. the solvent polarity parameter [ET(30)] for 6a (●x025CF;) and 6c (▲x025B2;). Source: Chen et al. [26], © 2012 The Royal Chemical Society.Figure 5.12 Schematic representation of the equilibrium between 6d and its aggregates in solution. At low a concentration and/or a high temperature, dimers of M‐type helicity were preferred; whereas, at high concentrations and/or low temperatures, extended polymeric stacks of P‐configuration were formed. Source: Dehm et al. [31], © 2007 American Chemical Society.Figure 5.13 Schematic representation of the solvent‐dependent self‐assembly/disassembly behavior observed for a binary mixture of 7 and an OPV derivative. Source: van Herrikhuyzen et al. [32], © 2004 American Chemical Society.Figure 5.14 Scanning electron microscopy images of the nanobelts and nanoparticles obtained from the self‐assembly of 9a and 10b, respectively. Source: Balakrishnan et al. [34]. Figure reproduced with kind permission; © 2006 American Chemical Society.Figure 5.15 Schematic representation of the two different stacking modes: Longitudinally displaced J‐type stacking of 11a (a) and rotationally displaced cofacial stacking of 11c (b). Source: Chen et al. [37]. © 2007 John Wiley and Sons.Figure 5.16 Schematic representation of amphiphilic squaramide dyes which gave supramolecular polymers due to dipolar π–π stacking interactions. Source: Bujosa et al. [49], © 2020 The Royal Chemical Society.Figure 5.17 Schematic representation of the boron‐centered triangulene 13 that self‐assembled into a columnar LC phase. Source: Kushida et al. [55]. © 2015 John Wiley and Sons.Figure 5.18 (a) Schematic representation of the different packing modes of bowl‐shaped molecules (from the non‐columnar type‐C assemblies only one representative example is shown). (b) Schematic representation of the P‐centered triangulenes 14. Source: Yamamura et al. [56], © 2016 American Chemical Society.Figure 5.19 Schematic representation of the competition between the homo‐ and hetero‐type stacking of donor–acceptor‐type molecules driven by π–π and charge–transfer stacking forces, respectively. Source: Modified from Han et al. [59].Figure 5.20 Schematic representation of the synthesis of an amphiphilic D–A‐type foldamer and its self‐assembly into a 1D nanostack. Source: Jalani et al. [64], © 2013 Royal Society of Chemistry.Figure 5.21 Schematic representation of the amphiphilic donor–acceptor molecule17 that could assemble into nanosheets or vesicles depending on its conformation. Source: Jalani et al. [65]. Licenced under CC BY 3.0.Figure 5.22 Schematic representation of the various types of molecular tweezers/guest recognition motifs. (a) DAD‐ and ADA‐type assemblies, (b) DADA‐ and ADADA‐type assemblies and (c) self‐complementary and head‐to‐tail dimerization. Source: Han et al. [59], © 2018 The Royal Chemical Society.Figure 5.23 Schematic representation of the metal‐induced activation (a) and deactivation (b) of molecular tweezers according to Petitjean et al. [71].Figure 5.24 Schematic representation of the formation of a π‐electron ion pair via ion metathesis. Source: Haketa and Maeda [72], © 2017 Royal Society of Chemistry.Figure 5.25 Schematic representation of the three different self‐assembly modes of π‐electron ion pairs: charge‐by‐charge (a), charge segregated (b), and intermediate stacking (c). Source: Dong and Maeda [74].Figure 5.26 Schematic representation of the formation of a π‐electron anion via the supramolecular complexation of a (halide) ion by a planar anion receptor. (a) cartoon representation and (b) at the example of 18. Source: Dong and Maeda [74], © 2013 Royal Society of Chemistry.Figure 5.27 Schematic representation of the formation of π‐electron anions by deprotonation of OH‐groups.Figure 5.28 Schematic representation of the π‐electron ion pairs, based on the cycloheptatrienyl (Ch+) cation and cyclopentadienyl anions (21). The representations show the packing mode of [Ch][21b] in the solid state (three different views are depicted). Source: Bando et al. [85], © 2016 John Wiley and Sons.Figure 5.29 (a) Schematic representation of the TATA+ cation 22 and of the receptor–anion complex [23⋅⋅⋅Cl]. (b) POM images, XRD patterns (at 101 °C) and proposed packing modes (from left to right) of the ion pairs [22][23⋅⋅⋅Cl] (a). Source: Haketa et al. [86]. Figure reproduced with kind permission; © 2010 Wiley‐VCHFigure 5.30 Schematic representation of the square‐planar Pt(II) complexes 24. The XRD pattern and the proposed packing mode for the ion‐pairing assembly of 24b and [23⋅⋅⋅Cl] is also shown. Source: Sekiya et al. [89], © 2014 Royal Society of Chemistry.Figure 5.31 Schematic representation of the molecular tweezers 25 and their reversible self‐assembly of 25a into a linear supramolecular polymer. Source: Kim et al. [90], © 2012 American Chemical Society.Figure 5.32 Schematic representation of the pH‐switchable molecular tweezers 26 that were used for the assembly of a packed liposome bilayer with 1,2‐dioctadecanoyl‐sn‐glycero‐3‐phosphocholine (DSPC) and N‐(carbonyl‐methoxypolyethyleneglycol‐2000)‐1,2‐distearoyl‐sn‐glycero‐3‐phosphoethanolamine (DSPE‐PEG2000). The packed liposome bilayer collapsed upon acidification. Source: Viricel et al. [94].© 2015 John Wiley and Sons.Figure 5.33 Schematic representation of the molecular tweezers27 containing two positively‐charged organoplatinum(II) pincers. Source: Tanaka et al. [95], © 2013 John Wiley and Sons.Figure 5.34 Schematic representation of the reversible supramolecular polymerization of 28. Source: Tian et al. [96], © 2014 John Wiley and Sons.Figure 5.35 Schematic representation of the complementary homoditopic monomers29 and 30 and their supramolecular polymerization in a DADA‐type fashion. The evolution of the specific viscosity as a function of the monomer concentration is also shown. Source: Tian et al. [97], © 2016 American Chemical Society.Figure 5.36 Schematic representation of the porphyrin‐based molecular tweezers 31 and of the self‐complementary dimerization of 31a. Source: Haino et al. [68], © 2006 American Chemical Society.Figure 5.37 Schematic representation of the supramolecular polymerization of 32. Source: Modified from Haino et al. [99].Figure 5.38 Schematic representation of the heteroditopic monomers33 and of the supramolecular polymerization of 33a. The specific viscosity of 33a in CHCl3 (white circles) and toluene (black circles) at 293 K is also depicted. Source: Haino et al. [100], © 2012 John Wiley and Sons.Figure 5.39 Schematic representation of the bis‐pyridine derivative34 that was used for the formation of a supramolecular polymer network in combination with 33b. A picture of the free‐standing polymer film of this material is also shown. Source: Hager et al. [102]. Figure reproduced with kind permission; © 2015 Wiley‐VCH.Figure 5.40 Schematic representation of the calix[4]pyrroles35 that adopted a 1,3‐alternate conformation, which could accommodate two guest molecules via CT interactions. The conformation of 35 could be switched reversibly by adding chloride anions. Source: Nielsen et al. [107], © 2004 American Chemical Society.Figure 5.41 Schematic representation of the supramolecular polymers derived from 35 and 36. A representative SEM image of the fiber‐like assemblies also shown. The polymers could be disassembled in the presence of competitive guest molecules or anions. Source: Park et al. [110]. Figure reproduced by kind permission; © 2011 National Academy of Science.Figure 5.42 Schematic representation of the supramolecular 2 : 1 complex formed by γ‐CD and C60 in water. Source: Based on Andersson et al. [114].Figure 5.43 Schematic representation of a molecular tweezers on the basis of a bis‐calix[5]arene (38) and its encapsulation of a C60 guest in the cavity. Source: Based on Haino et al. [116].Figure 5.44 Schematic representation of the bis‐exTTF‐based molecular tweezers39 and its solvent‐dependent interaction with C60. Source: Pérez et al. [118], © American Chemical Society.Figure 5.45 Schematic representation of the heteroditopic monomer40 and of the corresponding supramolecular CT polymer (40)n. The AFM images of (40)n deposited onto a mica surface are also shown (left: 2D image, 277 nm × 277 nm; right: 3D image, 80 × 65 nm). Source: Fernández et al. [119], Figure reproduced with kind permission; © 2008 Wiley‐VCH.Figure 5.46 Schematic representation of the preorganized exTTF‐based receptor 41 and the analogous AB‐type monomer 42. The MALDI‐TOF mass spectrum of (42)n is also depicted. Source: Isla et al. [120], © 2014 John Wiley and Sons.Figure 5.47 Schematic representation of the formation of a supramolecular hyperbranched polymer starting from an AB2‐type monomer. Source: Fernández et al. [121], © 2008 American Chemical Society.

6 Chapter 6Figure 6.1 Schematic representation of important crown ether hosts.Figure 6.2 Schematic representation of the (pseudo)rotaxane formation. Source: Zheng et al. [9], 2012 Royal Society of Chemistry.Figure 6.3 (a) Schematic representation of the AB‐type monomer 4 and the ball‐stick representation of the solid structure of the dimer (4)2 (counter ions were omitted for clarity). (b) MALDI time‐of‐flight mass spectrum of 4. Source: Cantrill et al. [35], © 2001 American Chemical Society.Figure 6.4 (a) Schematic representation of monomer 5. (b): Space‐filling representation of the [c2]daisy chain (5)2 (counterions were omitted for clarity). Source: Ashton et al. [38], © 1998 John Wiley and Sons.Figure 6.5 (a) Schematic representation of the self‐assembly of 6 into a polypseudorotaxane. (b) Schematic representation of the multifunctional monomer 7.Figure 6.6 (a) Schematic representation of the formation of a supramolecular polymer gel via the self‐assembly of a flexible AB‐type monomer. (b) Pictures of the supramolecular gel and its stimuli‐triggered gel–sol transitions (by changing either the temperature or pH value). Source: Dong et al. [41]. Figure reproduced with kind permission; © 2011 Wiley‐VCH.Figure 6.7 Schematic representation of the photoinduced supramolecular polymerization of 8 which was accompanied by sol–gel transition of the solution at high monomer concentrations. Source: Zheng et al. [43]. Figure reproduced with kind permission; © 2018 Elsevier B.V.Figure 6.8 Schematic representation of the self‐assembly of the complementary homoditopic monomers 9 and 10 (a) as well as 11 and 12 (b); the counterions were omitted for clarity [47]. Source: Figure reproduced with kind permission; © 2003 American Chemical Society.Figure 6.9 Schematic representation of polypseudorotaxane formation via selective crown ether recognition. Source: Wang et al. [49], © 2009 Royal Society of Chemistry.Figure 6.10 Schematic representation of the supramolecular self‐assembly of 15 and 12 (m = 10). The supramolecular polymer exhibited AIE‐based fluorescence in solution as well as in the solid state, i.e. in the electrospun nanofibers. Two representative SEM images (a/b) an a fluorescence microscopy image (c) of the fibers are shown. Source: Chen et al. [50], © 2015 The Royal Chemical Society.Figure 6.11 Schematic representation of the multifunctional monomer 16 and its transformation into a linear supramolecular polymer (poly‐16), covalent (co)polymers (17) and, eventually, covalent‐supramolecular polymers. A summary of the mechanical properties of these materials is also depicted. Source: Zhang et al. [51], © 2020 John Wiley and Sons.Figure 6.12 Schematic representation of the hetero‐tetratopic monomer 18, which gave a supramolecular polymer network due to a self‐cross‐linking process. Source: Wang et al. [52]. Figure reproduced with kind permission; © 2020 American Chemical Society.Figure 6.13 (a) Schematic representation of the formation of a host–guest assembly via threading MV2+ into the cavity of cryptand 19. (b) Schematic representation of the self‐assembly of AA‐type monomers with the BB‐type (MV2+)2‐monomer into linear supramolecular polymers. Source: Niu et al. [57], © 2011 American Chemical Society.Figure 6.14 Schematic representation of the formation of a cryptand‐based linear supramolecular polymer via three different reaction pathways. Source: Wei et al. [60].Figure 6.15 (a) Schematic representation of the unsymmetrical cryptand 20. (b) Schematic representation of the host–guest complex of 20 with the MV2+ and dibenzylammonium cation. (c) Schematic representation of the linear supramolecular polymer assembled from 20 and the ditopic guests 12 and 14. Source: Yao et al. [61], © 2018 John Wiley and Sons.Figure 6.16 Schematic representation of macromolecular building blocks for host–guest complexation, based on crown ether recognition. Source: Zheng et al. [9], © 2012 Royal Society of Chemistry.Figure 6.17 (a) Schematic representation of the (a) polymer chain extension via crown ether recognition. Source: Based on Gibson et al. [63]Figure 6.18 Schematic representation of the heteroditopic macromonomers 21. The nanofibers, fabricated from the chain‐extended polymer (21)n by electrospinning, showed a multiple degradation behavior. The representative SEM images show the fibers before (c) and after the respective treatment with K+ (a), NEt3 (b), Cl (d), and heat (e). Source: Chen et al. [123]. Figure reproduced with kind permission; © 2016 The Royal Chemical Society.Figure 6.19 Schematic representation of the self‐assembly of the AB2‐monomer 22 into a supramolecular hyperbranched polymer.Figure 6.20 Schematic representation of the formation of dendritic pseudorotaxanes via a convergent synthesis. Source: Gibson et al. [72], © 2002 American Chemical Society.Figure 6.21 Schematic representation of the formation of star‐shaped and graft polymers (a and b, respectively), based on crown ether recognition. Source: (a) Zheng et al. [9], © 2012 Royal Society of Chemistry; (b) Gibson et al. [75], © 2009 John Wiley and Sons.Figure 6.22 Schematic representation of the formation of dendrimers, based on crown ether recognition: (a) threading‐followed‐by‐stoppering and subsequent Wittig exchange. and (b) sliding methodology. Source: Elizarov et al. [80], © 2002 American Chemical Society.Figure 6.23 Schematic representation of the formation of a [2]rotaxane via a thermal azide to alkyne cycloaddition “click” reaction. Source: Ashton et al. [82], © 1996 John Wiley and Sons.Figure 6.24 (a) Schematic representation of the so‐called Wittig exchange – transformation of the “reactive” rotaxane I into the “inert” rotaxane III via intermediate II. Source: Rowan et al. [76], © 2000 American Chemical SocietyFigure 6.25 Schematic representation of the formation of a mechanically interlocked daisy chain using the Wittig exchange (a) and the “threading‐followed‐by‐swelling” methodology. Source: Rowan et al. [76], © 2000 American Chemical SocietyFigure 6.26 Schematic representation of a supramolecular muscle/shuttle, based on the pH‐switchable [c2]daisy chain (28)2. Source: Coutrot et al. [88], © 2008 American Chemical Society.Figure 6.27 Schematic representation of building block 23 and the derived poly([c2]daisy chain) that was utilized for expansion/contraction studies. Source: Clark et al. [93], © 2009 American Chemical Society.Figure 6.28 (a) Schematic representation of the synthesis of a poly[2]rotaxane via Sonogashira cross‐coupling polymerization. Source: Sasabe et al. [94]Figure 6.29 Schematic representation of the synthesis of dendritic [4]rotaxanes by “dynamic covalent chemistry” (DCC). Source: Leung et al. [99], © 2005 American Chemical Society.Figure 6.30 Schematic representation of main‐chain pseudorotaxane formation via two different synthetic strategies: Multiple threading of macrocycles onto a linear polymer chain (top) and polymerization of preassembled pseudorotaxanes, as supramonomers (bottom). Source: Wenz et al. [101], © 2006 American Chemical Society.Figure 6.31 Schematic representation of the synthesis of polyrotaxanes with stopper moieties within the polymers main chain.Figure 6.32 Schematic representation of the polystyrene‐based poly(pseudo)rotaxanes.Figure 6.33 Schematic representation of the hydrophilic macromolecular host 39 and the hydrophilic guest 40 that self‐assembled into a pH‐responsive supramolecular amphiphile; the amphiphile could be used as a nanocontainer for the encapsulation and controlled release of hydrophobic molecules, such as Nile red. Source: Ji et al. [121], © 2012 American Chemical Society.Figure 6.34 Schematic representation of the formation of an amphiphilic supramolecular diblock copolymer from 29 and 31 that assembled further into nanostructures depending on the length ratio of the constituent blocks (lsp: length of the supramolecular block, lcp: length of the covalent block). Source: Ji et al. [122], © 2013 John Wiley and Sons.

7 Chapter 7Figure 7.1 Schematic representation of the generalized CB[n] synthesis via co‐condensation of 1 with formaldehyde under acidic conditions. The dimensions of different CB[n] representatives are listed [5]. The ball‐stick representation of the solid‐state structure of CB[8] is also shown. Source: Kim et al. [4], © 2000 American Chemical Society.Figure 7.2 Schematic representation of the various exclusion and inclusion complexes of CB[n]s. Source: Barrow et al. [16].Figure 7.3 Schematic representation of the formation of supramolecular polymers via complexation of BB‐ or ABBA‐type monomers by a CB[n] host. Source: Redrawn from Yang et al. [33], © 2015 American Chemical Society.Figure 7.4 Schematic representation of the formation of a surface‐anchored supramolecular polymer (counterions omitted for clarity). Source: Kim et al. [53], © 2004 American Chemical Society.Figure 7.5 Schematic representation of the formation of the supramolecular polymer {4@CB[8]}n; the assembly into 1 : 1 or 2 : 2 species is prevented due to steric and electronic constraints, respectively. Source: Liu et al. [55], © 2010 John Wiley and Sons.Figure 7.6 Schematic representation of a supramolecular polymer based on metal‐to‐ligand coordination and host–guest complexation. Source: Liu et al. [56], © 2013 The Royal Chemical Society.Figure 7.7 (a) Schematic representation of the bifunctional monomer 6 (counterions were omitted for clarity). (b) Side view of the X‐ray single crystal structure of {6@CB[8]}n. Source: del Barrio et al. [57], © 2013 American Chemical Society.Figure 7.8 (a) Schematic representation of the ditopic monomers 7 to 9. (b) Schematic representation of the linear supramolecular polymer {9@CB[8]}n. Source: Liu et al. [59], © 2013 John Wiley and Sons.Figure 7.9 Schematic representation of the imidazolium‐containing guests 10 and 11. (a) Schematic representation of the step‐wise self‐assembly of 11 and CB[8] into linear polymer chains and, subsequently, fibrous aggregates. (b) Representative AFM images of the crystalline nanostructures obtained from 10/CB[8] (left) and 11/CB[8] (right). Source: Barrio et al. [62]. Figure reproduced with kind permission; © 2019 American Chemical Society. Licensed under CC BY 4.0.Figure 7.10 (a) Schematic representation of the self‐assembly of 12 and CB[8] into 2 : 2 dimers and linear polymers due to the formation of H‐ or J‐type aggregates within the hosts' cavities. (b) Representation of the structure of oligomeric {12@CB[8]}n according to molecular modeling. Source: Xu et al. [63], © 2011 The Royal Chemical Society.Figure 7.11 (a) Schematic representation of the homoditopic monomer 13. (b) Schematic representation of the proposed 2 : 1 binding mode. (c) AFM–SMFS curves of the supramolecular polymer {10@CB[8]}n and the Gaussian fitting thereof. Source: Tan et al. [60], © 2013 Royal Society of Chemistry.Figure 7.12 (a) Schematic representation of the formation of a cyclic 1 : 1 complex via a ring‐chain equilibrium pathway. (b) Schematic representation for the inhabitation of protein complexation due to the formation of the highly stable cyclic complex. Source: Ramaekers et al. [67], © 2013 Royal Society of Chemistry.Figure 7.13 (a) Schematic representation of the formation of a discrete 2 : 3 complex and (b) a 2D supramolecular organic framework due to radical dimerization stabilized by CB[8]. Source: Zhang et al. [72], © 2014 Royal Society of Chemistry.Figure 7.14 (a) Schematic representation of the self‐assembly of the MV˙+ radical cation and CB[8] into a 2 : 1 complex. (b) Schematic representation of the hexacationic ditopic guest 14. Source: Redrawn from ref. Yin et al. [76], © 2013 Chinese Chemical Society.Figure 7.15 Schematic representation of the proposed equilibrium for the self‐assembly of a bis‐CB[10] derivative with a bis‐adamantyl guest into discrete complexes and linear supramolecular polymers. Source: Nally and Isaacs [85], © 2009 Elsevier.Figure 7.16 Schematic representation of the synthesis of CyPnTD[n] macrocycles. The solid‐state structure of CyP4TD[4], as determined by single crystal X‐ray analysis, as well as the cavity dimensions of the CyPnTD[n] derivatives are also shown. Source: Wu et al. [86], © 2017 Royal Society of Chemistry.Figure 7.17 Schematic representation of the formation of a supramolecular polymer network with AIE behavior. A representative TEM image of the polymer particles is also depicted. Source: Wu et al. [88], © 2018 Royal Society of Chemistry. Figure reproduced with kind permission; © 2018 The Royal Chemical Society.Figure 7.18 Schematic representatio of the two‐step self‐assembly of the TPE dyes 15 and 16 into cubic or spherical nanostructures, respectively. Representative TEM images visualizing these nanostructures are also depicted. Source: Li et al. [90]. Figure reproduced with kind permission; © 2018 Wuiley‐VCH.Figure 7.19 Schematic representation of the self‐assembly of 2 : 1 heterotrimers due the formation of a CT complex within the CB[8] cavity; (a–c) three applications of this particular interaction using polymeric building blocks are also shown. Source: Rauwald and Scherman [96], © 2008 John Wiley and Sons.Figure 7.20 Schematic representation for the formation of a supramolecular ABA‐type triblock copolymer. Source: Zayed et al. [98], © 2014 Royal Society of Chemistry.Figure 7.21 Schematic representation of the step‐wise, self‐assembly of supramolecular micelles that could be used for the controlled release of doxorubicin, as a model drug. The release, as a function of the added reducing agent, is also shown. Source: Zhao et al. [99], © 2014 Royal Society of Chemistry.Figure 7.22 Schematic representation of the temperature‐triggered formation of double‐layered vesicles from supramolecular lipid–peptide conjugates. Source: Loh et al. [102], © 2014 Royal Society of Chemistry.Figure 7.23 Schematic representation of the stepwise formation of vesicles from an initial mixed micellar system by CB[8]‐triggered self‐sorting. Source: Mondal et al. [104], © 2014 American Chemical Society.

8 Chapter 8Figure 8.1 Schematic representation of the macrocyclic hosts calix[n]arene and resorcin[n]arene (R denotes any substituent; whereas, n and m = n − 3 refer to the number of phenyl moieties within the macrocycle).Figure 8.2 Schematic representation of the tetraurea‐functionalized calix[4]arene (a) and of its dimer with an encapsulated small guest molecule (b) [21]. Source: Redrawn from Dalcanale and Pinalli [19], © 2015 Springer Nature.Figure 8.3 Schematic representation of the self‐assembly of calix[4]arene dimers 2 into polycaps [21]. Source: Redrawn from Dalcanale and Pinalli [19], © 2015 Springer Nature.Figure 8.4 Schematic representation of the depolymerization of a polycap in the presence of a dimeric capsule into a dumbbell‐shaped 2 : 1 complex [21]. Source: Redrawn from Dalcanale and Pinalli [19]. © 2015 Springer Nature.Figure 8.5 Schematic representation of different polycaps derived from homo‐ or heteroditopic bis‐calixarenes (black calixarenes are equipped with aryl ureas, gray calixarenes carry sulfonyl ureas). Source: Castellano et al. [24], © 1998 American Chemical Society.Figure 8.6 (A) Photomicrographs of a typical Schlieren texture observed from a LC polycap in CHCl3 (a) and p‐difluorobenzene (b) as viewed between crossed polarizers. (B) Laser confocal microscopy images of fibers assembled from the LC phases of the polycap in CHCl3 either by sample shearing (a) or fiber pulling from the sample (b). Source: Castellano et al. [25]. Figure reproduced with kind permission; © 1999 Wiley‐VCH.Figure 8.7 Schematic representation of the bis‐calixarene 3 comprising a pH‐sensitive dipeptide linker that could be used to switch between a soluble and insoluble form of the resulting polycap. Source: Redrawn from Xu et al. [27], © 2003 American Chemical Society.Figure 8.8 Schematic representation of the two‐step assembly of a supercap by (a) calixarene dimerization in an apolar solvent and (b) binding of CO2. Source: Redrawn from Xu et al. [29].Figure 8.9 Schematic representation of the bis‐calixarene 5, as a building block for the two‐step assembly of a stimuli‐responsive organogel [30]. Source: Redrawn from Guo et al. [18], © 2012 Royal Society of Chemistry.Figure 8.10 Schematic representation of the bis‐calixarene 6 equipped with a pending pyrene moiety. Source: Modified from Xu and Rudkevich [32].Figure 8.11 Schematic representation of bis‐calixarenes, connected via their upper rims. Depending on the structure of linker inter‐ and/or intramolecular association is enabled. Source: Guo and Liu [18], © 2012 Royal Society of Chemistry.Figure 8.12 Schematic representation of the formation of a supramolecular polymer with vacant calixarene moieties. Source: Liu et al. [37], © 2002 Taylor & Francis.Figure 8.13 Schematic representation of the calix[4]arenes 8 and 9 equipped with H‐bonding moieties. These were used for the self‐assembly of supramolecular polymeric nanostructures. Source: Klok et al. [42], © 1999 American Chemical Society.Figure 8.14 Schematic representation of resorcin[4]arene 10 and calix[4]arene 11. Source: Baldini et al. [45], © 2011 American Chemical Society.Figure 8.15 Schematic representation of cavitands 12 and 13. The latter one was used as AB‐type monomer in a supramolecular polymerization. Source: Dalcanale and Pinalli [19], © 2015 Springer Nature.Figure 8.16 (a) Representation of the solid‐state structure of poly‐13 (perpendicular view to the polymer chains) [52]. (b) Schematic representation of the star‐shaped supramolecular polymer, derived from a template‐driven self‐assembly using a planar tetratopic guest, as core. Source: Yebeutchou et al. [52], © 2008 John Wiley and Sons.Figure 8.17 Schematic representation of the ptert‐butylcalix[5]arenes 14–16. Source: Pappalardo et al. [56].Figure 8.18 Schematic representation of the bis‐calix[5]arenes 17 and the alkyl‐diammonium salts 18 used for the self‐assembly into the supramolecular architectures A‐D. Source: Gattuso et al. [60], © 2008 American Chemical Society.Figure 8.19 Schematic representation of the homoditopic “bis‐container” 19 for the complexation of fullerene derivatives, such as the dumbbell‐shaped bis‐fullerene 20 and the polyacetylene 21 with fullerene side chains.Figure 8.20 Schematic representation of the formation of a supramolecular coil–rod–coil triblock copolymer. Source: Hirao et al. [67], © 2020 American Chemical Society.Figure 8.21 Schematic representation of the homoditopic AA‐type bis‐cavitand hosts (22) and BB‐type guests (23) used for supramolecular polymerization. Source: Tancini et al. [69], © 2010 John Wiley and Sons.Figure 8.22 Representation of the crystal structure of the linear supramolecular polymer obtained from the self‐assembly of 22c and MV2+ (the counterions are omitted for clarity). Source: Tancini et al. [69], © 2010 John Wiley and Sons.Figure 8.23 (a) Schematic representation of the p‐sulfonatocalix[n]arene 24 and the corresponding homoditopic hosts 25 and 26. (b) Schematic representation of the self‐assembly of 25 into a linear or net‐like polymer upon complexation of a dicationic or tetracationic guest. Source: Guo et al. [71], © 2009 John Wiley and Sons.Figure 8.24 Schematic representation of the tetracationic bis‐viologen guests 27 and their use in the self‐assembly with bis‐calixarene 26 to afford linear polymer or cyclic oligomers. Source:(a) Redrawn from Guo et al. [73,74] © 2010 Royal Society of Chemistry; (b) Redrawn from Qian et al. © 2012 John Wiley and Sons.Figure 8.25 Schematic representation of the synthesis of a ternary supramolecular polymer via a two‐step self‐assembly approach. Source: Redrawn from Qian et al. [74], © 2012 John Wiley and Sons.Figure 8.26 Schematic representation of the pH‐ and redox‐sensitive supramolecular polymer assembled from bis‐calixarene 26 and the heteroditopic guest 28. Source: Redrawn from Ma et al. [77], © 2011 Royal Society of Chemistry.Figure 8.27 Schematic representation of the multifunctional guest 29 and the chiral supramolecular polymer derived by binding of α‐CD and 26. The polymer showed a light‐driven isomerization that could be monitored by SEM imaging of the dried cast films on glass slides. Source: Sun et al. [78]. Figure reproduced with kind permission; © 2013 American Chemical Society.Figure 8.28 Schematic representation of calix[4]pyrrole (30). The four most relevant conformations (a) as well as the anion‐induced transformation from the 1,3‐alterante to the cone‐shaped conformation (b) are also shown. Source: Wu et al. [80], © 2001 The Royal Chemical Society.Figure 8.29 Schematic representation of the TTF‐functionalized calix[4]pyrroles 31, which, in their 1,3‐alternate conformation, accommodated two electron‐poor guests. The conformation of 33 could be switched reversibly by adding chloride anions. Source: Nielsen et al. [88], © 2004 American Chemical Society.Figure 8.30 Schematic representation of the dicationic calix[4]pyrrole‐based guests 32. (a) Schematic representation of the structure of the supramolecular polymer assembled from 31c and 32a; a representative SEM image visualizing the solid‐state morphology of the material is also shown. (b) Schematic representation of the supramolecular assemblies formed by the depolymerization of (31c⋅⋅⋅32a)n in the presence of TBAI (left) and TEAI (right). Source: Kim et al. [91], © 2013 American Chemical Society. Figure reproduced with kind permission; © 2013 American Chemical Society.Figure 8.31 (a) Schematic representation of the flexible, ditopic guests 33. (b) Evolution of the DP as a function of the monomer concentration in different solvents (MCH: methylcyclohexane; DCE: 1,2‐dichloroethane). Source: Bähring et al. [92], © 2014 The Royal Chemical Society.Figure 8.32 (a) Schematic representation of the complementary ditopic monomers 34 and 35, which gave a linear supramolecular polymer due to calix[4]pyrrole–carboxylate anion recognition. Source: Yuvayapan et al. [93], © 2019 The Royal Chemical SocietyFigure 8.33 Schematic representation of the PBI dye 30 that was assembled into fibers upon addition of a p‐sulfonatocalix[n]arene (24). The TEM, SEM, and AFM images (from the left to the right) of the nanostructures are also depicted. Source: Guo et al. [96]. Figure reproduced with kind permission; © 2012 The Royal Chemical Society.Figure 8.34 Schematic representation of the reversible self‐assembly of an amphiphilic molecule in the presence of a macrocyclic host. Source: García‐Rio and Basílio [101], © 2019 Elsevier.Figure 8.35 Schematic representation of the formation of a supramolecular amphiphile, as a thermoresponsive carrier for doxorubicin hydrochloride. A representative TEM image of the vesicles is also depicted. Source: Wang et al. [104]. Figure reproduced with kind permission; © 2010 Wiley‐VCH.Figure 8.36 Schematic representation of the multiple stimuli–responsive vesicles obtained from the self‐assembly of 24 (n = 4) and 39. Source: Wang et al. [105], © 2011 American Chemical Society.Figure 8.37 Schematic representation of the multistep assembly of a linear supramolecular polymer from the heterodifunctional guest 40 and two different types of hosts. The TEM images of the three types of assemblies observed in the study are also depicted. Source: Zhang et al. [106]. Figure reproduced with kind permission; © 2014 The Royal Chemical Society.Figure 8.38 Schematic representation of the self‐assembly of 41 into vesicles and, in the presence of Ag(I) ions, spherical micelles. Source: Redrawn from ref. Houmadi et al. [122], © 2007 American Chemical Society.Figure 8.39 Schematic representation of the homoditopic guest 42 that was assembled into spherical nanostructures in the presence of the sulfonated calix[4]arene 24 (a) or bis‐calix[4]arene 26 (b). Both types of nanostructures exhibited aggregation‐induced emission. Source: Redrawn from ref. Jiang et al. [107], © 2014 American Chemical Society.Figure 8.40 Schematic representation of calix[4]arene 43 as a host for the self‐assembly into supramolecular micelles with chlorin‐e6 (Ce6). The size distribution as determined by DLS measurements (a) and a TEM image of the micelles (b) are also depicted. Source: Tu et al. [111]. Figure reproduced with kind permission; © 2011 The Royal Chemical Society.Figure 8.41 Schematic representation of an enzyme‐responsive supramolecular amphiphile for drug‐delivery applications. Source: Guo et al. [113], © 2012 American Chemical Society.Figure 8.42 Schematic representation of the self‐assembly of 44 and 45 into supramolecular amphiphiles. Source: Redrawn from Kharlamov et al. [115], © 2013 American Chemical Society.Figure 8.43 Schematic representation of the self‐assembly of 46 and 47 into supramolecular nanosheets. The TEM (a) and AFM (b) images of these nanosheets are also depicted. Well‐defined nanosheets, which are 1–2 mm long and 300–500 nm wide, are observed. Source: Yi et al. [116]. Figure reproduced with kind permission; © 2012 The Royal Chemical SocietyFigure 8.44 Schematic representation of a calix[8]arene‐based polypseudorotaxane. Source: Yamagishi et al. [119], © 2001 American Chemical Society.

9 Chapter 9Figure 9.1 Schematic representation of basic cyclodextrin (CD) structures. Source: Szejtli [4]. © 1997 American Chemical Society.Figure 9.2 (a) Schematic representations of a pseudorotaxane and a rotaxane. (b) Qualitative representation of the energy scheme for the disassembly of a [2] rotaxane into its molecular components. Source: Wenz et al. [11]. © 2006 American Chemical Society.Figure 9.3 Schematic representation of the synthesis of polypseudorotaxanes via two different approaches: multiple threading onto a preformed polymer and polymerization of a preformed pseudorotaxane monomer (a so‐called supramonomer). Source: Redrawn from Wenz [11]. © 2006 American Chemical Society.Figure 9.4 (a) Schematic representation of the two possible highly ordered modes for the packing of CD macrocycles along a polymer chain. (b) Schematic representation of a single‐ and double‐stranded assembly. Source: Redrawn from Ref. [11]. © 2006 American Chemical Society.Figure 9.5 (a) Solid‐state structure of an oligo(ethylene glycol) with threaded α‐CD rings and (b) scanning‐microscopy image of PEG with threaded α‐CD rings. (c) Plot of the threading kinetics of PEG and α‐CD (turbidity was measured by UV/vis absorption spectroscopy; the arrow indicates the end of the proposed induction period). Source: (a) Harada et al. [18]. © 2007 The Royal Chemical Society, (b) Arunachalam and Gibson [41]. © 2009 The Royal Chemical Society and © 1997 American Chemical Society, (c) Ceccato et al. [43]. © 1997 American Chemical Society.Figure 9.6 Schematic representation of the stimuli‐responsive self‐assembly of an azobenzene‐PEG‐CD conjugate. Source: Inoue et al. [45]. © 2007 The Royal Chemical Society.Figure 9.7 Schematic representation of the transformation of a randomly coiled polymer into a highly ordered crystalline state. Source: Tonelli [62]. © 2009 Springer Nature.Figure 9.8 Schematic representation of the polypseudorotaxane formation and isolation, the coalescence process, and the coalesced polymer. Source: Tonelli [62]. © 2009 Springer Nature.Figure 9.9 Schematic representation of the polydimethylsilane‐based polypseudorotaxane featuring a helix‐type conformation within the supramolecular assembly. Source: Redrawn from Ref. [76]. © 2003 The Royal Chemical Society.Figure 9.10 Schematic representation of a polypseudorotaxane formed by SWCNTs and η‐CD. Source: Dodziuk et al. [82]. © 2003 The Royal Chemical Society.Figure 9.11 Schematic representation of poly(bolaamphiphile)s 4–6 used as macromolecular guests for polypseudorotaxane formation (the counterions are omitted for clarity). Source: Herrmann et al. [22]. © 1997 American Chemical Society.Figure 9.12 (a) Schematic representation of the self‐assembly of components 6 and 7 into a flexible poly[2]rotaxane; (b) possible modes of binding of Gal‐1 to this supramolecular receptor. Source: Nelson et al. [84]. © 2004 American Chemical Society.Figure 9.13 Schematic representation of polypseudorotaxanes, based on block copolymers, star‐shaped polymers, and graft/comb polymers. Source: Redrawn from Wenz et al. [11]. © 2006 American Chemical Society.Figure 9.14 Atomic‐force microscopy (AFM) images of the platelets formed by the aggregation of pluronics‐based polypseudorotaxane [(a) phase image, (b) topography image, scale bar: 1 μm]. Source: Perry et al. [91]. Figure reproduced with kind permission; © 2011 The Royal Chemical Society.Figure 9.15 Schematic representation of the triblock copolymer 8 used for the site‐selective complexation by α‐CD. The resultant rod‐coil‐rod polymer featured a crystallization‐driven self‐assembly that could be modulated via the concentration and/or the pH value (two different morphologies, i.e. nanorings and nanorods, observed at different pH values are also shown). Source: Qi et al. [103]. © 2015 John Wiley and Sons.Figure 9.16 Schematic representation of the site‐selective complexation of a diblock copolymer by β‐CD moieties. A representative TEM image of the micelles assembled from the resultant amphiphilic pseudorotaxane‐PS diblock copolymer in water is also shown. Source: Cho and Allcock [114]. Figure reproduced with kind permission; © 2009 American Chemical Society.Figure 9.17 Schematic representation of the synthetic routes toward (a) polyrotaxanes with terminal stopper moieties, (b) polyrotaxanes with stopper moieties within the main chain, and (c) polyrotaxanes with stopper moieties attached to the main chain. Source: Wenz et al. [11]. © 2006 American Chemical Society.Figure 9.18 Schematic representation of a polyrotaxane formation facilitated by supramolecular host–guest chemistry. Source: Herrmann et al. [123]. © 1997 John Wiley and Sons.Figure 9.19 Representative scanning‐tunneling microscopy (STM) images of the polyazamethine‐based polyrotaxane on a highly ordered pyrolytic graphite (HOPG) surface (a) 2D representation, (b) 3D representation, (c) Molecular model of the structure. Source: (a, b) Liu et al. [126]. Figure reproduced with kind permission. © 2004 American Chemical Society.Figure 9.20 Schematic representation of the three‐step synthesis of polyrotaxanes, featuring π‐conjugated backbones. Source: Michels et al. [25]. © 2003 John Wiley and Sons.Figure 9.21 Schematic representation of the synthesis of a poly[2]rotaxane with a π‐conjugated polymer backbone. Source: Redrawn from Terao [133]. © 2011 The Royal Chemical Society..Figure 9.22 (a) Schematic representation of a chain‐extended polyrotaxane formed by a photochemical dimerization of the attached 2‐anthryl end groups. (b) Schematic representation of the reversible formation of linear chain‐extended polyrotaxanes and polycatenanes from a polyrotaxane. Source: (a) Okada et al. [136]. © 2004 American Chemical Society, (b) Okada et al. [137]. © 2003 American Chemical Society.Figure 9.23 Schematic representation of the supramolecular polymerization using (a) a heteroditopic AB‐type CD–guest conjugate, (b) an equimolar mixture of homoditopic CD–CD dimer and guest dimer (AA‐ and BB‐type monomers, respectively), and (c) a CD of sufficient cavity size and a BB‐type guest dimer. Source: Redrawn Yang et al. [138]. 2015 American Chemical Society.Figure 9.24 Schematic representation of the host–guest conjugates 12 and the supramolecular polymer poly‐12a. Source: Harada et al. [140]. © 2001 Springer Nature.Figure 9.25 Schematic representation of the (a) double‐threaded dimer and (b) daisy‐chain trimer, obtained from the self‐assembly of CDC<span class="dbond"></span>;C–cinnamoyl conjugates. Source: Hoshino et al. [142]. © 2000 The Royal Chemical Society.Figure 9.26 (a) Schematic representation of the host–guest conjugate 13 and its self‐assembly into a helical supramolecular polymer. (b) STM image of poly‐13 deposited onto a MoS2 substrate. Source: (a) Miyauchi et al. [145]. © 2005 American Chemical Society.Figure 9.27 Schematic representation of the AB‐type monomer 14 for which supramolecular polymerization was enabled; whereas, the competing pathways were disabled. A representation of the SANS data, which corroborated a cylindrical structure of poly‐14 with a length and diameter of c. 6.6 and 1 nm, respectively, is also shown. Source: Evenou et al. [146]. © 2018 John Wiley and Sons.Figure 9.28 Schematic representation of the self‐assembly of 15 and 16 into an alternating supramolecular copolymer. Source: Miyauchi and Harada [147]. © 2004 American Chemical Society.Figure 9.29 Schematic representation of the synthesis of a poly[2]rotaxane from the self‐assembly of the AB‐type [2]rotaxane monomer 17. Source: Miyauchi et al. [148]. © 2005 American Chemical Society.Figure 9.30 Schematic representation of the homoditopic AA‐ and BB‐type components 18 and 19, respectively, utilized for the supramolecular polymerization. Source: Ohga et al. [149]. © 2005 American Chemical Society.Figure 9.31 Schematic representation of the formation of a supramolecular polymer using the monomers γ‐CD and 20. The supramolecular polymer could reversibly be converted into a covalent polymer, upon light irradiation. Source: From Zhang et al. [153]. © 2013 American Chemical Society.Figure 9.32 Schematic representation of the formation of a POM‐containing polyrotaxane by a supramolecular polymerization and subsequent light‐driven anthracene dimerization. Representative TEM images showing fibrous bundles and isolated polymer chains are also depicted. Source: Guan et al. [154]. Figure reproduced with kind permission; © 2019 The Royal Chemical Society.Figure 9.33 Schematic representation of the formation of a supramolecular diblock copolymer, based on a CD–adamantyl host–guest complex [PNIPAM: poly(N‐isopropyl acrylamide), PMeOx: poly(2‐methyl‐2‐oxazoline)]. Source: Redrawn from Ref. [172]. © 2011 American Chemical Society.Figure 9.34 Schematic representation of the assembly of core–corona micelles from the polystyrene–β‐CD conjugate. Source: Giacomelli et al. [168]. © 2009 American Chemical Society.Figure 9.35 Schematic representation of the amphiphilic supramolecular diblock copolymer, which exhibited a redox‐responsive behavior due to the selectivity of β‐CD to bind Fc and to release Fc+. The TEM images show the reversibility of the voltage‐driven disassembly/reassembly process. Source: Yan et al. [169]. © 2010 The Royal Chemical Society.Figure 9.36 Schematic representation of the formation of light‐responsive nanotubes, assembled from an amphiphilic supramolecular diblock copolymer, for the controlled release of rhodamine‐B. The release rate, as a function of the irradiation time, is also shown. Source: Yan et al. [171]. Figure reproduced with kind permission; © 2010 American Chemical Society.

10 Chapter 10Figure 10.1 Schematic representation of pillar[5]arene (1) and two cationic guests (2 and 3) that formed host–guest complexes of 1 : 1 stoichiometry with 1. Source: Ogoshi et al. [1]. © 2008 American Chemical Society.Figure 10.2 Schematic representation of the viologen oligomers 4 used as macromolecular guests.Figure 10.3 Schematic representation of the pillar[5]arene‐based pseudorotaxane of PANI's emeraldine base and its reduction to the leucoemeraldine species.Figure 10.4 Schematic representation of the chain extension of polyethylene induced by the threading of 1 onto the chains. Source: Ogoshi et al. [29]. © 2013 Springer Nature.Figure 10.5 (a) Schematic representation of the pillar[5]arene‐based polymer‐containing rotaxanes 5. (b) Representation of the Ru‐catalyzed cross‐linking of 5b into a topological polymer gel. Source: Ogoshi et al. [31]. © 2014 The Royal Chemical Society.Figure 10.6 Schematic representation of the host–guest conjugate 6 and its reversible formation of a self‐inclusion complex. Source: Guan et al. [33]. © 2011 The Royal Chemical Society.Figure 10.7 Schematic representation of the monofunctional pillar[5]arenes 7 that assembled into supramolecular dimers (7a and 7b) or a supramolecular polymer (7c). (a) Ball‐and‐stick model of such a supramolecular dimer. (b) Representative SEM image of a gold‐coated fiber of (7c)n. Source: Liu et al. [35]. © 2012 American Chemical Society.Figure 10.8 Schematic representation of the host–guest conjugate 8. (a) Image of the supramolecular gel (8)n; (b) optical microscopy image (500×) of aggregates of (8)n; (c) SEM image of the gel in CH2Cl2. Source: Strutt et al. [38]. © 2012 The Royal Chemical Society.Figure 10.9 Schematic representation of the photo‐switchable host–guest conjugate 9 and its self‐assembly into supramolecular architectures. Source: Wang et al. [45]. © 2014 The Royal Chemical Society.Figure 10.10 Schematic representation of the host–guest conjugate 10 that assembled into a linear supramolecular polymer and, subsequently into aggregates via π–π stacking interactions. Source: Fathalla et al. [46]. © 2015 The Royal Chemical Society.Figure 10.11 Schematic representation of the BODIPY‐containing host (11) and guest molecules (12 and 13). For the supramolecular polymer (1112)n, FRET from the donor dye to the acceptor one was observed. Source: Meng et al. [49]. Figure reproduced with kind permission; © 2015 The Royal Chemical Society.Figure 10.12 Schematic representation of the two‐step self‐assembly of the pillarene‐based building blocks 14 and 15 into a supramolecular polymer. Source: Wang et al. [51]. © 2013 The Royal Chemical Society.Figure 10.13 Schematic representation of the fluorescent host 16 and guests 17 which were employed, as monomers, to assemble linear supramolecular polymers. The device configuration, containing these materials, as emissive layer (EML), as well as the obtained electroluminescence spectra are also depicted. Source: Yang et al. [52]. © 2017 American Chemical Society.Figure 10.14 Schematic representation of the formation of a dynamic polyrotaxane via a two‐step self‐assembly process. A TEM microscopy image of the supramolecular polymer obtained from 1,4‐diaminobutane (n = 4) is also shown. Source: Hu et al. [54]. Figure reproduced with kind permission; © 2012 American Chemical Society.Figure 10.15 Schematic representation of the two‐step self‐assembly of a polypseudorotaxane containing X‐bonding and host–guest interactions within the main and the side chains, respectively. The solid‐state structure of the assembly, as determined by XRD analysis is depicted. A comparison of the diffusion coefficients (D), as determined by DOSY measurements as well as a SEM image of the materials are also shown. Source: Liu et al. [55]. Figure reproduced with kind permission; © 2018 Wiley‐VCH.Figure 10.16 Schematic representation of the formation of a double‐dynamic polymer, based on supramolecular host–guest interactions and dynamic covalent chemistry. Source: Xu et al. [56]. © 2013 American Chemical Society.Figure 10.17 Schematic representation of a supramolecular polymer, based on an exo‐wall and endo‐cavity complexation. Source: Wang et al. [59]. © 2015 The Royal Chemical Society.Figure 10.18 Schematic representation of the synthesis of columnar oligo‐pillarenes according to Stoddard et al. (n = 0–9). Source: Strutt et al. [60]. © 2014 John Wiley and Sons.Figure 10.19 Schematic representation of the self‐assembly of 22 into a hyperbranched supramolecular polymer. The distribution of the hydrodynamic diameter (Rh) according to DLS as well as a representative TEM image of the obtained nanoparticles are also shown. Source: Xiaoyang et al. [61]. Figure reproduced with kind permission; © 2013 John Wiley and Sons.Figure 10.20 (A) Schematic representation of the formation of a dynamic supramolecular polymer network. (B) The TEM images of the polymer at certain concentrations are also shown [(a–c) 1, 2, and 5 mM]. Source: Zhang et al. [62]. Figure reproduced with kind permission; © 2013 The Royal Chemical Society.Figure 10.21 Schematic representation of the two‐step self‐assembly of a cross‐linked supramolecular polymer. Source: Hu et al. [69]. © 2013 The Royal Chemical Society.Figure 10.22 Schematic representation of the PPE‐type polymer 25 and an anion‐responsive polypseudorotaxane prepared thereof. Source: Sun et al. [71]. Figure reproduced with kind permission; © 2013 The Royal Chemical Society.Figure 10.23 Schematic representation of the amphiphilic pillar[5]arene 26 and the nanostructures assembled thereof. Source: Yao et al. [73]. © 2012 American Chemical Society.Figure 10.24 Schematic representation of the amphiphilic pillar[5]arene 27 and its reversible self‐assembly into a gel‐like material or reverse giant vesicles. Source: Gao et al. [78]. Figure reproduced with kind permission; © 2013 The Royal Chemical Society.Figure 10.25 Schematic representation of the reversible vesicle formation via supramolecular interactions between pillar[6]arene 29 and the photo‐switchable guest 28. Source: Yu et al. [81]. Figure reproduced with kind permission; © 2012 American Chemical Society.Figure 10.26 Schematic representation of drug‐loaded vesicles exhibiting a light‐induced release. Source: Hu et al. [82]. © 2015 John Wiley and Sons.Figure 10.27 Schematic representation of the self‐assembly of 30 and 31 into vesicular nanoparticles; the TEM images show representative nanoparticles observed in both cases (scale bar: 200 nm). Source: Cao et al. [86]. Figure reproduced with kind permission; © 2015 American Chemical Society.

11 Chapter 11Figure 11.1 The important combinations of orthogonal supramolecular interactions. Source: Li et al. [17]. © 2012 The Royal Chemical Society.Figure 11.2 (a) Schematic representation of the metallo‐supramolecular polymer 1; (b) performance of 1 in an asymmetric hydrogenation reaction compared with Rh(I)‐MonoPhos, as the benchmark. Source: Yu et al. [9]. © 2010 John Wiley and Sons.Figure 11.3 Schematic representation of the bis‐terpyridine ligand 2 that was used for the formation of the homometallic and heterobimetallic metallopolymers comprising Fe(II) and/or Ru(II) ions. The polymers exhibited pronounced characteristics in their photophysical properties as revealed by their coloration that could be switched electrochemically by applying an external voltage. Source: Hu et al. [37]. Figure reproduced with kind permission; © 2013 The Royal Chemical Society.Figure 11.4 Schematic representation of the heteroditopic macroligands3 and 4, as monomers for the preparation of multi‐stimuli‐responsive supramolecular polymers of high molar mass. Source: Hofmeier et al. [39]. © 2005 American Chemical Society.Figure 11.5 Schematic representation of the self‐assembly of 5 and 6 in presence of Fe(II) ions. Source: Grimm et al. [45]. © 2011 John Wiley and Sons.Figure 11.6 Schematic representation of supramolecular block copolymers 7–9, based on metal‐to‐ligand coordination and H‐bonding interactions. Source: Mansfeld et al. [48]. © 2013 The Royal Chemical Society.Figure 11.7 Schematic representation of the self‐assembly of 10 and Pd(II) ions into a double‐cross‐linked polymer. Source: Chen et al. [50]. © 2019 The Royal Chemical Society.Figure 11.8 (a) Schematic representation of the formation of a double‐stranded helical supramolecular polymer via metal‐to‐ligand coordination and ionic interactions. (b) Representation of the size distribution of the polymer (data derived from DLS measurements). Source: Ikeda et al. [51]. © 2006 American Chemical Society.Figure 11.9 Schematic representation of the heteroditopic building block 12 and its use in the formation of a multi‐stimuli‐responsive supramolecular polymer. Source: Gröger et al. [52]. © 2011 American Chemical Society.Figure 11.10 Schematic representation of a supramolecular pseudorotaxane with Pd(II) complexes in the main chain. Source: Zhu et al. [54]. © 2011 American Chemical Society.Figure 11.11 (a) Schematic representation of the two‐step, self‐assembly of 15 into a cross‐linked supramolecular polymer; (b) representative SEM images of (15)n with increasing amount of Pd(II) ions (top to bottom: 0%, 20%, 60%, and 100%, respectively). Source: (a) Yan et al. [59]. © 2011 John Wiley and Sons, (b) Yan et al. [59]. Figure reproduced with kind permission; © 2012 Wiley‐VCH.Figure 11.12 Schematic representation of the supramolecular gel formed by the cross‐linking of metallocages via crown ether recognition. The material exhibited a multi‐stimuli‐responsive behavior (e.g. switchability between a gel‐like and liquid state). Source: Lu et al. [60]. Figure reproduced with kind permission; © 2018 American Chemical Society.Figure 11.13 Schematic representation of the formation of the 1 : 1 complex 16@CB[6] and its metallo‐supramolecular polymerization into linear chains or a 2D polymer networks when Cu(II) or Ag(I) ions, respectively, are added. Source: Whang et al. [62]. © 1996 American Chemical Society.Figure 11.14 Schematic representation of the two‐step assembly of 1617 with CB[6] and Cd(II) ions into a double‐chained supramolecular 1D polymer. Source: Redrawn from Park et al. [63].Figure 11.15 Schematic representation of the typ ligands 18–20, as monomers, for the two‐step supramolecular polymerization: (a) homometallic polymer, based on homotrimeric host–guest complexes, and (b) heterobimetallic polymer, based on heterotrimeric host–guest complexes. Source: (a) Liu et al. [70]. © 2013 The Royal Chemical Society, (b) Joseph et al. [71]. © 2014 American Chemical Society.Figure 11.16 Schematic representation of supramolecular triblock copolymer 21. Source: Yang et al. [73]. © 2010 American Chemical Society.Figure 11.17 Schematic representation of the formation of 1D‐stacks from 22 by synergistical H‐bonding interactions. Source: Nieuwenhuizen et al. [75]. © 2010 John Wiley and Sons.Figure 11.18 Schematic representation of homoditopic polymer 23. The AFM imaging revealed fiber formation as a result from substituents in 6‐position (scale bar: 100 nm). Source: Appel et al. [76]. Figure reproduced with kind permission; © 2011 American Chemical Society.Figure 11.19 Schematic representation of polymers 24–26 and their self‐assembly behavior. Source: Mes et al. [79]. © 2011 American Chemical Society.Figure 11.20 Schematic representation of the supramolecular architecture 27. Source: Shao et al. [81]. © 2004 American Chemical Society.Figure 11.21 Schematic representation of monomers 28 and 29 as well as their self‐assembly into a linear supramolecular polymer; the SEM image of a fiber grown from (28⋅⋅⋅29)n is also shown. Source: Li et al. [85]. Figure reproduced with kind permission; © 2011 The Royal Chemical Society.Figure 11.22 Schematic representation of the self‐assembly of 30 and 31 into a pseudorotaxane network. Source: Li et al. [87]. © 2011 The Royal Chemical Society.Figure 11.23 (a) Schematic representation of the self‐assembly of a star‐shaped molecule 32 into a helical columnar triplex; (b) schematic representation of the fabrication of an ordered array of columnar lines on a mica substrate. Source: van Hameren et al. [88]. © 2008 American Chemical Society.Figure 11.24 (a) Schematic representation of OPE 33 and (b) its concentration‐dependent self‐assembly into nanoparticles, microspheres or giant fibrillar structures. Source: Ajayaghosh et al. [90]. © 2006 John Wiley and Sons.Figure 11.25 Schematic representation of the two‐step, self‐assembly of 34 and 35 into a supramolecular superlattice; the polarizing optical microscope (POM) image is also shown (34‐to‐35 ratio of 8 : 1, 100 °C). Source: Fitié et al. [92]. © 2008 American Chemical Society.Figure 11.26 Schematic representation of the self‐assembly of 36 into supramolecular polymers and grafted polymers (in the absence and presence of 37, respectively). Representative TEM images of the spherical micelles and 1D fibers, obtained from the coassembly of 36 with 37a and 37b, respectively, are also shown. Source: Jamadar and Das [94]. © 2020 The Royal Chemical Society.Figure 11.27 Schematic representation of the hierarchical self‐assembly of the sulfonamide‐functionalized oligopeptide 38. Source: Teng et al. [96]. © 2019 John Wiley and Sons.Figure 11.28 Schematic representation of the supramolecular polymerization of 39 and 40 due to orthogonal host–guest complexation. Source: Wang et al. [98]. © 2008 American Chemical Society.Figure 11.29 Schematic representation of the polypseudorotaxane formation via an orthogonal three‐component self‐assembly process. Source: Wang et al. [103]. © 2009 The Royal Chemical Society.Figure 11.30 Schematic representation of the self‐assembly of a polycationic dendron with naphthalene dicarboxylic acids into cylindrical and spherical aggregates. Source: Redrawn from Gröhn et al. [104]. © 2019 John Wiley and Sons.Figure 11.31 Schematic representation of the pair of oppositely charged porphyrins (412− and 424+/5+); a representative TEM image of the resultant nanotubes is also shown (inset: a tube trapped in a vertical orientation by a thick mat of tubes). Source: Wang et al. [109]. Figure reproduced with kind permission; © 2004 American Chemical Society.Figure 11.32 Schematic representation of the oppositely charged PBI (43) and phthalocyanine dyes (44), which gave a supramolecular helically stacked superstructure. Source: Guan et al. [110]. © 2005 John Wiley and Sons.Figure 11.33 Schematic representation of the proposed model of the two‐step, self‐assembly of the carboxy‐functionalized NBI 45 and the tetra‐cationic guanidiniocarbonyl pyrrole 46. Source: Samanta et al. [113]. © 2005 John Wiley and Sons.Figure 11.34 Schematic representation of the functionalized monomers 47 and 48. (a) Schematic representation of the supramolecular polymer formed by 47, CB[8], and 48 in a directional two‐step assembly approach. (b) The AF4 elution curve of the supramolecular polymer (water, as eluent, UV detector). Source: Song et al. [114]. © 2014 The Royal Chemical Society.Figure 11.35 Schematic representation of the heteroditopic monomers 49–51 used for the sequence‐controlled supramolecular copolymerization. Source: Hirao et al. [117]. © 2017 Springer Nature.Figure 11.36 Schematic representation of the formation of a supramolecular polymer due to CT interactions and pillar[5]arene‐based host–guest complexation. Source: Chen et al. [118]. © 2016 The Royal Chemical Society.Figure 11.37 (a) Schematic representation of the self‐assembly of the a myoglobin‐dimer and streptavidin with heme‐bis(biotin) conjugate 52; (b) AFM height image of the supramolecular polymer on a modified mica substrate (100 mM K3PO4 buffer, pH value of 7); (c) SEC traces (UV/vis absorption detector) of the supramolecular polymer (solid line) and the fragment obtained after cleavage of the disulfide bridges (dashed line). Source: Oohara et al. [24]. Figure reproduced with kind permission; © 2012 Wiley‐VCH.Figure 11.38 (a) Schematic representation of the formation of the supramolecular polymers (53)n, based on protein recognition. (b) SEC analysis of (53)n using reference proteins, as standards. Source: Kitagishi et al. [123]. © 2007 American Chemical Society.Figure 11.39 Schematic representation of the reversible switching between a thermodynamically and a kinetically stable architecture (linear polymer chain vs. spherical micelle). Source: Oohora et al. [124]. © 2017 The Royal Chemical Society.

12 Chapter 12Figure 12.1 Theoretical plot of the degree-of-polymerization (DP) vs. association constant (Ka in M−1) for a supramolecular polymerization (two different concentrations, isodesmic self‐assembly model). Source: Brunsveld et al. [2]. © 2001 American Chemical Society.Figure 12.2 Representative plots of the probability of binding (p) according to Weber [24] as a function of the total guest concentration ([G]0) and the Ka value (left axis) as well as the dissociation constant (Kd, right axis) at three different host concentrations [(a)–(c): [H]0 = 10−3 M, 10−5 M, and 10−7 M]. The thick horizontal bar indicates where [H]0 is equal to Kd and Ka−1. The contour lines represent 0.1 units of p (the p range from 0.2 to 0.8 is shaded gray. Source: Thordarson [20]. © 2011 Royal Society of Chemistry.Figure 12.3 Simulated binding isotherms (1 : 1 complexation, Ka = 100 M−1) for [H]0 = 10−6 M (red curve) and 10−5 M (blue curve); a random error of 2% was included in the simulation. Source: Thordarson [20]. © 2011 Royal Society of Chemistry.Figure 12.4 (a) SEC traces of [1]n and polystyrene references (eluent: CHCl3); (b) SEC traces of 2 and {[Ru(2)](BF4)2}n (eluent: DMAc, containing 5.5 mM NH4PF6). Source: (a) Ogawa and Kobuke [35]. © 2000 John Wiley & Sons, (b) Meier et al. [32]. © 2006 American Chemical Society.Figure 12.5 Schematic representation of the synthesis of the poly(daisy chain) 3. The SEC traces of 3 and of the utilized monomer are also shown (eluent: DMF containing 0.2 M NH4PF6). Source: Fang et al. [42]. © 2009 American Chemical Society.Figure 12.6 (a) Schematic representation of the formation of the supramolecular polymers (4)n. (b) Reduced viscosity of (4a)n in CH2Cl2 (▲x025B2;) and CCl4 (●x025CF;), the corresponding calculated curves are also shown (dashed and solid line, respectively). Source: Abed et al. [48]. © 2001 Elsevier.Figure 12.7 Schematic representation of the ring‐chain‐mediated polymerization of 5 and 6. The bilogrithmic viscosity–concentration plot allowed the determination of the cpc value. Source: Huang et al. [50]. © 2007 American Chemical Society.Figure 12.8 Schematic representation of the supramolecular polymerization of monomer 7 in the absence (a) and presence of the chain‐stopping agent 8 (b). Source: Lortie et al. [56]. © 2005 American Chemical Society.Figure 12.9 Schematic representation of the β‐cyclodextrin‐dimer (9) and the adamantyl‐dimer (10). The evolution of the Mn of the supramolecular polymer (9···10)n as a function of the β‐CD‐dimer concentration (determined by VPO) is also shown. Source: Hasegawa et al. [62]. © 2005 American Chemical Society.Figure 12.10 (a) Schematic representation of the stepwise synthesis of a water‐soluble polyrotaxane. (b) Plot of the experimental Mw according to AUC as a function of the theoretical DP (DPnom, the nonideal behavior at high DP values was attributed to accidental chain termination and fractionation upon purification. Source: From Michels [87]. © 2003 John Wiley & Sons.Figure 12.11 Comparison of the 1H NMR spectra of the metallo‐supramolecular polymer 11 and the corresponding metal‐free ligand (both recorded in DMSO‐d6). In the spectrum of 11, the signals assigned to the end groups are asterisked. Source: Brunsveld et al. [89]. © 2015 American Chemical Society.Figure 12.12 Schematic representation of the supramolecular polymerization of 12a. The concentration dependency of the diffusion coefficient (D) is also shown. Source: Haino et al. [92]. © 2012 John Wiley & Sons.Figure 12.13 (A) Schematic representation of the chain‐extended supramolecular polymer {[(13)Fe(13)](PF6)2}n. (B) DOSY of {[(13)Fe(13)](PF6)2}n in the absence (a) and presence of TFA (b). Source: Mansfeld et al. [97]. © 2013 The Royal Society of Chemistry.Figure 12.14 Analysis of a H‐bonded supramolecular polymer in the keto‐ and the enol‐form by 1H DQ‐MAS NMR spectroscopy. Source: Schnell et al. [99]. © 2002 The Royal Chemical Society.Figure 12.15 Positive turbo ion spray time‐of‐flight MS of 14. Source: Miyauchi et al. [63]. © 2005 American Chemical Society.Figure 12.16 (a) Schematic representation of the self‐assembly of 15 and C60 into a supramolecular polymer; (b) STM image of the polymer deposited onto a HOPG surface (the inset shows the corresponding line profile; (c) schematic representation of the polymer with structural parameters. Source: (a, c) Liu et al. [114]. © 2006 American Chemical Society, (b) Figure reproduced with kind permission; © 2006 American Chemical Society.Figure 12.17 (a, b) Schematic representation of the supramolecular polymer 16 and the proposed binding motif. (c) STM image of a monolayer of 16 at the liquid–graphite interface (self‐assembled from a 1,3,5‐trichlorobenzene solution, the scale bar is 4 nm). Source: (a, b) Ciesielski et al. [117]. Figure reproduced with kind permission; © 2009 John Wiley & Sons, (c). Ciesielski et al. [117]. © 2009 Wiley‐VCH.Figure 12.18 Schematic representation of the formation of loops within metallo‐supramolecular polymers. The corresponding UHV‐LT‐STM images are also depicted. Source: Li et al. [120]. © 2020 American Chemical Society.Figure 12.19 (a) Schematic representation of a metallo‐supramolecular Sierpiński hexagonal gasket. (b) Visualization of the assembly by UHV‐LT‐STM imaging. Source: Newkome et al. [121]. Figure reproduced with kind permission; © 2006 AAAS.Figure 12.20 (a) Schematic representation of two metallo‐supramolecular “spiderwebs.” (b) The corresponding UHV‐LT‐STM images are also depicted. Source: Wang et al. [122]. Figure reproduced with kind permission; © 2019 American Chemical Society.Figure 12.21 AFM images of spin‐coated solutions of 17 (a–c) and 18 (d) on mica substrates; for 17, various ligand‐to‐metal ratios were utilized (a: 0.6, b: 0.85, c: 0.98). Source: Schwarz et al. [126]. Figure reproduced with kind permission, © 2010 American Chemical Society.Figure 12.22 (a) Schematic representation of the experimental setup; (b) histogram of the bound rapture forces at a pulling velocity of 118 nm s−1. Source: Kudera et al. [138]. © 2003 John Wiley & Sons.Figure 12.23 Schematic representation of the self‐assembly of 19 into a linear supramolecular polymer, based on two different types of noncovalent interactions; the cryo‐TEM images of {[Fe(19)2]Cl2}n are also shown (a: no staining, b: negative staining with uranyl acetate). Source: Gröger et al. [143]. Figure reproduced with kind permission; © 2011 American Chemical Society.Figure 12.24 (a) Schematic representation of the formation of a “polynanocage”; (b and c) AFM and TEM image of the polynanocage. Source: (a) Fan et al. [149]. © 2012 American Chemical Society; (b, c) Fan et al. [149]. Figure reproduced with kind permission; © 2012 American Chemical Society.Figure 12.25 Schematic representation of the formation of a pseudorotaxane of high molar mass that could be processed into rod‐like fibers. Source: Wang et al. [152]. Figure reproduced with kind permission; © 2009 Royal Society of Chemistry.Figure 12.26 Schematic representation of supramolecular diblock copolymers, based on multiple H‐bonding interactions and their temperature‐dependent microphase‐separation behavior [the two different representations denote a polyisobutylene and a poly(ether ketone) constituent block, respectively]. Source: Binder and Zirbs [145]. © 2006 Springer.Figure 12.27 SAXS data for the metallo‐supramolecular polymer obtained from macroligand 20 with Zn(II) ions [SAXS curves at various 20 to Zn(II) ratios are shown]. Source: Burnworth et al. [13]. © 2011 Nature Publishing Group.Figure 12.28 Representation of the solid‐state structure of the β‐CD–PEG polypseudorotaxane, with the PEG chain reaching through four unit cells of β‐CD dimers. The H‐bonding interactions between pairs of head‐to‐head connected CDs are shown (view along the b‐axis, hydrogen atoms are omitted for clarity). Source: Udachin et al. [164]. © 2000 American Chemical Society.Figure 12.29 (a) Schematic representation of the supramolecular polymer formed by the polyassociation of the AB‐type monomer 21. (b) Two views of the 1D arrangement of (21)n (only the hydrogen atoms involved in the C–H···π interactions are shown, all other hydrogen atoms were omitted for clarity). Source: Zhang et al. [169]. © 2011 John Wiley & Sons.Figure 12.30 The form factor (P) for two polymers whose chains follow self‐avoiding walk statistics and which have a Rg of 100 Å. The Guinier and intermediate regions can be clearly seen. For (a) the chains are perfectly monodisperse (Ð of 1.0) and for (b) the Ð value is large (i.e. Mw/Mn = 6.4). The dashed line (c) qualitatively indicates the effect of a persistence length (L) of c. 10 Å, giving a third region with a slope of −1. Source: Whittell et al. [173]. © 2013 John Wiley & Sons.Figure 12.31 (a) Schematic representation of the metallo‐supramolecular polymerization of 21 with various divalent metal ions [e.g. Co(II), Ni(II)]. (b) The Co(II)‐containing polymer showed an electrochemically driven switchability between the gel and sol states. Source: (a) Gasnier et al. [179]. © 2009 American Chemical Society, (b) Gasnier et al. [179]. Figure reproduced with kind permission; © 2009 American Chemical Society.Figure 12.32 (a) Comparison of the molar mass and size detection/separation range for different techniques; (b) Picture of a typical AF4 setup (www.schubert-group.uni-jena.de); (c) Schematic representation of the separation field in a AF4 channel (www.nanolytics.de). Source: (a) Yang et al. [177]. © 2015 American Chemical Society, (b) U.S. Schubert, (c) © Nanolytics.Figure 12.33 (a) Schematic representation of the supramolecular self‐sorting polymerization of 24 with two different cucurbituril derivatives. (b) Representation of the ratio‐dependent AF4 elution curves obtained by the MALS detector (the labeling of the curves indicates the applied CB[8]‐to‐CB[7] ratios). Source: Huang et al. [184]. © John Wiley & Sons.Figure 12.34 Schematic representation of the formation of single‐chain nanoparticles by intramolecular metallo‐supramolecular crosslinking of ligand‐functionalized polymers. Source: Neumann et al. [185].Figure 12.35 (a) Idealized representation of a TDA measurement in which the particles travel from the injection point to the detector with a residence time (t0); in this way, the particles are dispersed by a combination of pressure‐driven laminar flow and translational diffusion. (b) TDA profile of three different SCNPs. Source: (a) Lemal et al. [188]. © 2019 American Chemical Society, (b) Neumann [185].

Supramolecular Polymers and Assemblies

Подняться наверх