Supramolecular Chirality in Self-Assembled Systems - Chemical

Jul 20, 2015 - (2) The concept of molecular chirality has long been recognized and provided guidance in the design of drugs and functional molecules, ...
3 downloads 20 Views 105MB Size
Review pubs.acs.org/CR

Supramolecular Chirality in Self-Assembled Systems Minghua Liu,* Li Zhang, and Tianyu Wang Beijing National Laboratory for Molecular Science (BNLMS), CAS Key Laboratory of Colloid, Interface and Chemical Thermodynamics, Institute of Chemistry, Chinese Academy of Sciences, Beijing 100190, People’s Republic of China 4.3.6. Sonication 4.3.7. pH Value 4.4. Chiral Amplification in Supramolecular Systems 4.4.1. Analogue-Induced Chiral Amplification 4.4.2. Chiral Amplification in Binary Systems 4.4.3. Chiral Amplification to Nanoscale 4.4.4. Unexpected Amplification in Racemate Assemblies 4.5. Chiral Memory in Supramolecular Systems 4.5.1. Helicity Memory in Noncovalently-Induced Helical Polymers 4.5.2. Chiral Memory in Aggregates Such as J and H Aggregates 4.5.3. Helicity Memory in Chiral Cages from Coordination Compounds 5. Spontaneous Symmetry Breaking and Emergence of Supramolecular Chirality in Self-Assembled Systems from Exclusively Achiral Molecules 5.1. Liquid-Crystal and Banana-Shaped Molecules 5.2. Solution Systems, Micelles 5.3. Gel Systems 5.4. Air/Water Interface and LB Films 5.5. Controlling Handedness of Supramolecular Chirality 5.5.1. Vortices and Spin Coating 5.5.2. Circularly-Polarized Light 5.5.3. Surface Pressure 5.6. Self-Assembly of Racemic Systems 6. Applications of Supramolecular Chirality 6.1. Supramolecular Chiral Recognition and Sensing 6.2. Supramolecular Chiroptical Switches 6.3. Supramolecular Chiral Catalysis 6.4. Optics and Electronics Based on Supramolecular Chiral Assembly 6.5. Circularly Polarized Luminescence (CPL) Based on Chiral Supramolecular Assemblies 6.6. Biological Applications of Supramolecular Chirality 7. Conclusions Author Information Corresponding Author Notes Biographies

CONTENTS 1. Introduction 2. Basic Concepts Related to Molecular and Supramolecular Chirality 2.1. Configuration and Conformation Chirality 2.2. Induced Chirality 2.3. Helicity or Helical Chirality 3. Characterization of Supramolecular Chirality 3.1. Morphology Observation 3.2. Spectroscopic Methods for Characterization of Chirality 3.2.1. CD Spectra of Supramolecular Systems 3.2.2. Measurement Aspects 3.2.3. CD Spectra and Interpretation 4. Supramolecular Chirality in Self-Assembled Systems Containing Chiral Molecular Components 4.1. Supramolecular Chirality in Assemblies of Chiral Components 4.1.1. Amphiphiles 4.1.2. C3-Symmetric Molecules 4.1.3. π-Conjugated Molecules 4.1.4. Molecules with Multiple Chiral Centers 4.2. Chirality Transfer in Systems Containing Chiral and Achiral Molecules 4.2.1. Chirality Transfer through Noncovalent Bonds 4.2.2. Chirality Transfer from Solvent to Assemblies 4.2.3. Chirality Transfer from Low Molecular Weight Molecules to Macromolecules 4.3. Dynamic Features and Regulation of Supramolecular Chirality 4.3.1. Solvents 4.3.2. Temperature 4.3.3. Redox Effect Chirality 4.3.4. Photoirradiation 4.3.5. Chemical Additives © 2015 American Chemical Society

7305 7305 7306 7306 7307 7308 7308 7309 7309 7309 7309 7310 7310 7310 7313 7316 7324 7325 7325 7329 7330 7331 7332 7335 7336 7336 7336

7339 7340 7340 7340 7343 7343 7344 7344 7344 7346 7347

7347 7348 7349 7353 7354 7358 7358 7361 7362 7362 7365 7366 7372 7374 7381 7381 7383 7384 7384 7384 7384 7384

Special Issue: 2015 Supramolecular Chemistry Received: December 8, 2014 Published: July 20, 2015 7304

DOI: 10.1021/cr500671p Chem. Rev. 2015, 115, 7304−7397

Chemical Reviews Acknowledgments References

Review

biological systems and self-assembly, and many assist in developing new drugs and materials. In this review, we present an overview of the progress in supramolecular chirality in self-assembly systems, which mainly include self-assembly in solution or in dispersion systems, supramolecular gels, organized molecular films such as Langmuir and Langmuir−Blodgett films, and others. Although there are several reviews detailing supramolecular chirality, self-assembly, as well as chiral nanomaterials and nanostructures,7−13 the field has grown rapidly, and many new exciting results and phenomena have emerged. Furthermore, a general view of the chirality issue through the prism of supramolecular chirality will be helpful in better understanding many emergent chiral phenomena. In this review, we try to provide a comprehensive understanding of various organized chiral systems from the perspective of supramolecular chirality, with reference mainly to work published after 2010. First, we will provide a general overview of the supramolecular chirality, its definition, and special features through a comparison with the molecular chirality. Second, we will simply introduce the various modern techniques of characterization used in supramolecular chirality. Third, we will show in relative detail how molecular chirality could be transferred or related to supramolecular chirality in selfassembled systems containing chiral molecular components. Here, we will further show some special features of supramolecular chirality such as dynamic chirality, the principles governing the chiral amplification, and chiral memory. In the fourth part, we will discuss how achiral molecules can selfassemble into a chiral system, i.e., symmetry breaking and the emergence of supramolecular chirality in systems containing exclusively achiral molecules. A great challenge in the supramolecular chiral systems constructed from achiral molecules is the control of the chirality of system. Thus, we will discuss the manner of controlling the supramolecular chirality in systems composed of achiral molecules. Finally, we will show some typical applications of supramolecular chiral systems in electrooptics, sensing, asymmetric catalysis, biological applications, among others. In this portion, we will focus on the uniqueness of chiral supramolecular systems, how they differ from molecular chiral systems, and the new properties that emerge from supramolecular chirality. Currently, with the rapid development of supramolecular chemistry, self-assembly, and nanoscience, chirality has become an important issues, and many new chirality-related topics have appeared, such as chirality at a surface,14−16 chirality in a coordination system,17−22 and plasmonic chirality.23 These topics have been discussed and reviewed but are beyond the scope of this review.

7385 7385

1. INTRODUCTION Chirality is a basic characteristic of living matter and nature. During the evolution of life on our planet, nature has favored one kind of chirality, thereby selecting the L-amino acids (with the exception of glycine) as the main component of proteins and enzymes and D-sugars as the main components of DNA and RNA. In addition, chirality is universal and can be observed at various hierarchical levels from subatomic and molecular to supramolecular, nanoscopic, macroscopic, and galactic scales.1 Figure 1 illustrates some typical chiral substances and objects at these various scales. At a subatomic level, chirality is connected to parity conservation. Therefore, only left-handed helical neutrinos are found. At a molecular level, there are a huge number of chiral molecules in natural system such as amino acids, sugars, and terpenes, and many synthetic compounds are also chiral. Furthermore, there are many biological macromolecules or supramolecular systems with chirality, microorganisms with helix-shaped viruses, and bacteria such as tobacco mosaic virus and Helicobacter pylori, respectively, and macroscopic living systems such as snails. On a larger scale, one finds that many plants express chiral sense, such as mountain climbing vines. On a light-year scale, our galaxy system is also chiral. Among these various levels, chirality at a molecular and supramolecular level is of vital importance since it is strongly related to chemistry, physics, biology, materials, and nanoscience, which treat the matter in scales from atomic to molecular and supramolecular.2 The concept of molecular chirality has long been recognized and provided guidance in the design of drugs and functional molecules, while chirality at a supramolecular level is currently attracting great attention due to rapid developments in supramolecular chemistry and molecular self-assembly. Supramolecular chemistry is the chemistry beyond molecules or the chemistry of entities generated by intermolecular noncovalent interactions.3,4 Supramolecular chemistry is strongly related to self-assembly, which has been defined as the autonomous organization of components into patterns or structures without human intervention.5 Both molecular selfassembly and supramolecular chemistry are connected by noncovalent bonds and/or certain nano/microsized architectures. Molecular self-assembly plays an important role in biological systems, the transfer and storage of genetic information in nucleic acids, and the folding of proteins into efficient molecular machines.6 During such biological processes, supramolecular chirality, which can be simply regarded as chirality at a supramolecular level, is the result of biological molecular self-assembly. A typical example is the secondary structures of proteins, which can exhibit various conformations such as α-helix, β-sheet, and random coil structures with different supramolecular chirality. During the molecular self-assembly, supramolecular chirality is also the result of the special spatial arrangements of the molecules. Although supramolecular chirality is strongly related to the chirality of the component chiral molecules, it is not necessary that all components be chiral. To this end, achiral molecules can also possibly produce supramolecular chirality in a self-assembled system. Therefore, a deeper exploration of chirality at the molecular and supramolecular level will provide a better understanding of

2. BASIC CONCEPTS RELATED TO MOLECULAR AND SUPRAMOLECULAR CHIRALITY Chirality is used to describe an object that cannot be superimposed on its mirror image. When a molecule is not superimposable on its mirror image, then the molecule can be termed a chiral molecule. However, in practice, when judging whether a molecule is chiral, it is preferable to see if there is an asymmetric carbon atom in the molecule. An asymmetric carbon atom or chiral carbon is a sp3 carbon atom that is attached to four different types of atoms or four different groups of atoms. In addition, if a molecule possesses two noncoplanar rings that are dissymmetrically connected and cannot easily rotate about the chemical bond connecting them or the molecule possesses an axis about which a set of substituents is held in a spatial 7305

DOI: 10.1021/cr500671p Chem. Rev. 2015, 115, 7304−7397

Chemical Reviews

Review

Figure 1. Chiral architectures at various scales, from neutrinos to enantiomeric molecules, nanosized biomacromolecules with chiral structures (DNA and proteins), self-assembled micrometer-sized helical ribbons, microorganisms (helix-shaped bacteria), macroscopic living systems (seashells and plants), and galaxies. SEM image showing a helix is reprinted with permission from ref 285. Copyright 2014 John Wiley & Sons. The picture of a protein structure was obtained from Wikipedia (http://upload.wikimedia.org/wikipedia/commons/f/f3/T7RNA_polymerase_at_work.png) and reprinted under the “fair use” under Wikipedia’s license. Pictures are obtained from the following Web sites and apply to “fair use”: bacteria (http://tech.sina.com. cn/d/2010-01-29/10113816919.shtml), seashell (http://news.hainan.net/hainan/yaowen/tupian/2014/08/14/2016639.shtml), and flower (http:// www.chla.com.cn/htm/2011/0403/80069.html). The picture of a galaxy is a free stock graphic obtained from http://www.rgbstock.com/bigphoto/ mVErmjU%2FSpiral+Galaxy.

arrangement that is not superimposable on its mirror image, the molecule can also be chiral even if it lacks an asymmetric carbon atom. Such chirality is termed planar chirality and axial chirality, respectively. Thus, molecular chirality can be essentially classified as point, plane, and axis chirality. Since supramolecular chemistry is based on the chemistry of a noncovalent bond, supramolecular chirality is produced by nonsymmetric arrangement of molecules through a noncovalent bond. Therefore, supramolecular chirality can be produced from chiral component molecules, the combination of chiral and achiral molecules, or exclusively achiral molecules. Supramolecular chirality is largely dependent on the manner of assembly of the molecular components, but the chirality of the component molecule plays an important role in determining this manner in supramolecular systems. Generally, chiral molecules tend to form specific chiral structures with determined supramolecular chirality. In the combination of chiral and achiral molecules, achiral molecules can be induced into chiral assemblies if there is a strong interaction between the chiral and the achiral molecules. In most cases, the supramolecular chirality of the system is also determined and may follow the chirality of the chiral molecules. In the case of exclusively achiral components, supramolecular chirality can result from the formation of supramolecular systems but in general will be racemic in the resulting macroscopic system. Table 1 lists a simple comparison between molecular and supramolecular chirality. Both share some common features and are strongly related. When we speak of supramolecular chirality, molecular chirality should be frequently considered. For example, in the case of peptides, the chiral monomers covalently

polymerize into polymers to form chiral primary structures and then self-assemble into secondary and tertiary structures through noncovalent bonds, where both molecular and supramolecular chirality are involved. The key to their difference originates from the differences in the covalent and noncovalent bonds. There are some unique features of supramolecular chirality, as shown in Table 1. For example, supramolecular chirality is generally dynamic and changes in response to external stimuli and the environment. Chiral memory effects can also be seen in many supramolecular systems. Molecular chirality can originate from the tetrahedral geometry of certain atoms or the asymmetric axes and planes, while supramolecular chirality can be due to selfassembled helical, spiral structures and chiral sheets or chiral domain structures on surfaces. It should be noted that herein we mean molecular chirality generally refers to that of small molecules. If we consider the chirality of polymer systems, the distinctions between this and supramolecular chirality are less obvious. For example, the sergeant−soldier rule and the majority rule of chirality were originally proposed based on polymers and are also applicable to supramolecular systems. Below are some general terms related to molecular and supramolecular chirality. 2.1. Configuration and Conformation Chirality

Configuration refers to the permanent geometry resulting from the spatial arrangement of a system’s bonds. Conformation refers to the spatial arrangement of substituent groups that are free to assume different positions in space without breaking any bonds, because of the freedom of bond rotation. While configuration chirality is generally used in the case of molecular chirality, such as the absolute configuration of a chiral molecule, conformation chirality usually refers to supramolecular chirality in systems such as the secondary and tertiary structures of proteins. However, this term has not always been rigorously used based on this definition.

Table 1. Comparison between Molecular and Supramolecular Chirality molecular chirality composition bond chiral geometry manifestation of chirality naming convention special feature

atom covalent bond tetrahedron, axis, plane point, axis, and plane

supramolecular chirality

2.2. Induced Chirality

molecule, building block, tacton noncovalent bond helical, spiral, chiral sheet, chiral domain

R/S, L/D, M/P

conformation, secondary and tertiary structures, helicity, induced chirality, etc. M/P

fixed chirality, recognition

dynamic, sergeant−soldier rule, majority rule, chiral memory, recognition

Induced chirality generally refers to those chiral supramolecular systems where chirality is induced in an achiral guest molecule as a result of asymmetric information transfer from a chiral host or vice versa. This host could be a chiral molecule, chiral pocket, cavity, or chiral nanostructure. In order to produce the induced chirality, it is necessary for the achiral molecule to have a strong interaction with the chiral host through a noncovalent bond. A typical example of induced chirality is the encapsulation of a chromophore into the cavity of cyclodextrin.24 7306

DOI: 10.1021/cr500671p Chem. Rev. 2015, 115, 7304−7397

Chemical Reviews

Review

Figure 2. Some typical chiral molecules and their corresponding naming conventions.

Figure 3. (Top) Comparison of various microscopies used to characterize the chiral architectures. (Bottom) (A) AFM images of xerogels self-assembled from L- and D-HDGA (N,N-hexadecanedioyl-diglutamic acid). Reprinted with permission from ref 27. Copyright 2010 Royal Society of Chemistry. (B) STM images of the chiral twin chains from PVBA. Reprinted with permission from ref 28. Copyright 2001 The American Physical Society. (C) Mirrorimaged nanorods self-assembled from TPPS and (1R,2R)- or (1S,2S)-1,2-diaminocyclohexane. Reprinted with permission from ref 29. Copyright 2013 Royal Society of Chemistry. (D) TEM image of a chiral twist self-assembled from pyridine-containing L-glutamide. Reprinted with permission from ref 30. Copyright 2011 Royal Society of Chemistry.

2.3. Helicity or Helical Chirality

molecule with axial chirality. If the substituents are molecules

Helicity is a special form of axial chirality, which is defined as an entity that has an axis about which a set of substituents is held in a spatial arrangement that is not superimposable on its mirror image. If these substituents are atoms of molecular groups covalently attached to the axis then it can be classified as a chiral

held together along the axis by noncovalent bonds then the assemblies can be regarded as helical and have helical chirality. Helicity is very common in the supramolecular systems, and in particular, such chirality can often be visualized through AFM, 7307

DOI: 10.1021/cr500671p Chem. Rev. 2015, 115, 7304−7397

Chemical Reviews

Review

SEM, and TEM observations. Helicity can be classified as M or P helicity, which will be discussed later. Besides these chirality concepts, the naming conventions of chirality are complicated, and the detailed descriptions have been published.25−27 For the reader’s convenience, Figure 2 illustrates some typical chiral compounds or assemblies with certain types of chiral conventions. The R/S system is the most important and general nomenclature system for denoting enantiomers. Hereby, each chiral center is labeled R or S according to a system where its substituents are each assigned a priority, according to the Cahn− Ingold−Prelog priority rules (CIP). The D/L system (coined from the Latin dexter and laevus, right and left) is related to glyceraldehyde, whose two chiral isomers are labeled D and L. In this system, compounds are named by analogy to glyceraldehyde. Many biological molecules are labeled using this method. The M/P chirality generally refers to a supramolecular system or the axial or planar molecular chirality. Viewing from either end of a molecule or supermolecule downward along the helical axis, the system has P helicity if the rotation is clockwise and M helicity if the rotation is anticlockwise. The Λ and Δ chirality terms are used for defining coordination compounds. The enantiomers can be designated as Λ for a left-handed twist of the propeller described by the ligands and Δ for a right-handed twist, as illustrated in Figure 2.

Atomic force microscopy (AFM) is based on measurement of the force between a sharp tip and a sample’s surface. The sample is mounted on a piezoelectric scanner that moves the sample beneath a tip mounted on a soft cantilever. As the sample passes beneath the tip, the force between the tip and the surface can be measured, which forms an AFM image. AFM has been used successfully to probe the surfaces at scales down to the atomic level in vacuum, air, or other environments. The sample is generally fabricated on a very flat surface such as those of silica or mica. For example, AFM was used to observe self-assembled chiral nanotubes obtained through the gelation of bolaamphiphiles terminated with L- or D-glutamic acids.27 On the basis of the AFM observation, we can directly judge the supramolecular chirality of the nanotube. L-HDGA formed a right-handed helical nanotube, while D-HDGA formed a left-handed one. Scanning tunneling microscopy (STM) technology is a technology based on quantum tunneling. When a conducting tip is brought very close to a conductive surface and a bias voltage is applied, a tunneling current flows between the tip and the surface. The resulting tunneling current is a function of the gap between the tip and the surface.31 If the tunneling current is monitored and kept constant by adjusting the gap, the elevation of the surface can be traced and thus displayed an STM image. This technique provides an excellent means for controlling the distance between the probe and the surface and a very high resolution image of the samples mounted on an atomically flat conductive substrate such as HOPG. The STM technique provides a molecular level resolution and is used to directly discriminate the absolute configuration of chiral molecules.32 Further, the technique is also applicable to observation of the supramolecular chirality at surfaces.14 For example, Figure 3B shows mirror-imaged chiral twin chains that were self-assembled from PVBA (4-[trans-2-(pyrid-4-vinyl)]benzoic acid) adsorbed on a palladium substrate. The twin chains display supramolecular chirality.28 The scanning electron microscope (SEM) is the most widely used electron microscope for investigating the surface features of materials. When electrons interact with atoms in the sample they produce various signals that can be detected including scattered electrons and X-rays. SEM uses electron illumination to form images from the reflected electrons. SEM is critical in all fields that require characterization of solid materials. The SEM image is seen in three dimensions, but the result is a two-dimensional photograph; thus, it is especially useful for detecting chiral structures such as helices or twists. Figure 3C shows our results with self-assembled twisted nanorods by treating water-soluble TPPS with (1R,2R)- or (1S,2S)-1,2-diaminocyclohexane in which the mirror-imaged left-handed and right-handed helices were formed, respectively.29 Transmission electron microscopy (TEM) is a microscopy technique in which a beam of electrons is transmitted through an ultrathin specimen, interacting with the specimen as it passes through. A TEM image is formed from the interaction of the electrons transmitted through a very thin specimen ( 5 mm), C = crystallization. Reprinted with permission from ref 321. Copyright 2010 John Wiley & Sons.

The gelation properties, supramolecular chirality, and nanostructures of the racemic hydrogels can be regulated by changing molar ratios of different molecular building blocks (Figure 115).323 In addition to forming supramolecular gels, the mixing of enantiomers has also been used to tune the properties of different supramolecular assemblies. Oda et al. studied the twists and nanotubes formed from the coassembly of nonchiral dicationic n2-n Gemini amphiphiles with chiral tartrate anions. They found that the morphologies of the assemblies, such as the twist pitch of the ribbons, can be continuously modulated by varying the enantiomeric excess of tartrate anions. For instance, adding 10 mol % of the opposite enantiomer of tartrate anions led to a 15%

self-assembly can be a powerful tool for developing novel functional soft matters (Figure 114).322 Wang and Liu et al. studied the coassembly of the glutamicacid-based bolaamphiphile racemates with melamine. In this system, the coassembly of melamine with pure enantiomeric glutamic-acid-based bolaamphiphile (HDGA, 15b) cannot form gels. The assembly of the glutamic-acid-based bolaamphiphile racemate produced only precipitates. Mixing the glutamic-acidbased bolaamphiphile racemate with melamine produced good supramolecular gels. Remarkably, the racemic hydrogels showed a lower CGC value, enhanced mechanical rigidity, and dual pHresponsive ability compared to the pure enantiomer hydrogels. 7364

DOI: 10.1021/cr500671p Chem. Rev. 2015, 115, 7304−7397

Chemical Reviews

Review

Figure 115. (A) Molecular structure of the glutamic-acid-based bolaamphiphile (HDGA, 15b). (B) Schematic presentation of the racemic hydrogels formed by the coassembly of HDGA racemate and melamine. (C) Gelation properties of the (L + D)-HDGA mixtures and the (L + D)-HDGA/melamine mixtures with different ee values. Reprinted with permission from ref 323. Copyright 2014 American Chemical Society.

increase in the diameter of assembled supramolecular nanotubes (Figure 116).69 Liu et al. synthesized enantiomeric L- or D-glutamic-acid-based lipids (1) and investigated the self-assembly of these chiral enantiomers. Although both L- and D-enantiomeric molecules self-assembled into ultralong nanotubes, mixing D- and Lenantiomers with different molar ratios further changed the nanostructures consecutively from helical nanotubes to nanotwists to flat nanoplates (Figure 117).53 In most cases, the self-assembly of chiral molecules can lead to helical nanostructures, while racemates usually assemble into flat nanostructures. However, this situation has exceptions. Liu et al. synthesized the enantiomeric L- or D-alanine derivatives (AlaC17, 100), and the self-assembly and gelation properties of the corresponding individual enantiomers and the racemates were investigated. The self-assembly of individual enantiomers of AlaC17 can form only flat nanostructures, even though both LAlaC17 and D-AlaC17 can form gels in different organic solvents. Interestingly, racemic AlaC17 was found to self-assemble into beautiful twisted ribbons. Moreover, these twists are very sensitive to a slight enantiomeric excess of many other amino acids, showing remarkable macroscopic chirality. Therefore, these racemic assemblies can be used for the discrimination of various amino acid derivatives (Figure 118).230

Figure 116. (A) Coassembly of nonchiral dicationic n-2-n Gemini amphiphiles with chiral tartrate anions. (B) Transition from twisted ribbons to helical ribbons and then to tubules observed by TEM measurement. Reprinted with permission from ref 69. Copyright 2007 American Chemical Society.

Figure 117. (A) Correlative plot of the vibration bands of N−H, amide I, and amide II and the d spacing of the nanostructures in the mixed gels against the enantiomeric excess value (ED/L). (B) Proposed mechanism for the formation of various nanostructures upon mixing enantiomeric Lor D-glutamic-acid-based lipids. Reprinted with permission from ref 53. Copyright 2010 John Wiley & Sons.

6. APPLICATIONS OF SUPRAMOLECULAR CHIRALITY Recently, self-assembled chiral supramolecular systems have attracted greater attention due to the many potential applications of forms of soft matter. Such applications can further enhance our understanding of supramolecular chirality. In particular, new properties that single chiral molecules do not have emerge from supramolecular chiral systems. Certainly, for many applications of functional soft matters, supramolecular chirality plays the critical role, which is also dependent on the properties and characteristics of the supramolecular chiral information ex-

pressed on different materials within diverse scales.11,324−326 Herein, we address the most prominent fields for the application of supramolecular chirality, such as chiral recognition, sensing, and catalysis. Some optical devices based on supramolecular 7365

DOI: 10.1021/cr500671p Chem. Rev. 2015, 115, 7304−7397

Chemical Reviews

Review

Figure 118. (A) Molecular structure of the enantiomeric alanine derivatives AlaC17 (100). (B) Schematic illustration of the molecular packing for a single enantiomer and the racemate. The backgrounds are the SEM images of the transparent hexane gel formed by L-AlaC17 and the twisted ribbons formed by the racemic mixture. (C) Intensity of the CD signals (centered at 309 nm) of the racemate upon addition of 2 mol % of various amino acid derivatives. LBG is a glutamic acid derivative with two long alkyl chains. Reprinted with permission from ref 230. Copyright 2013 John Wiley & Sons.

be demonstrated by NMR, MS, light-scattering, and calorimetric measurements.

chirality and biological applications of chiral soft matters will also be discussed. 6.1. Supramolecular Chiral Recognition and Sensing

Chiral recognition is a very important issue in supramolecular chemistry. In general, interactions between different enantiomers and another chiral molecule can produce totally different results. These different interactions can be detected by spectral measurements or determination of their crystal structures,318,327−329 representing the chiral recognition. Thus far, studies on host−guest chemistry related to chirality have focused on chiral recognition,330−337 and many excellent works have been published. For example, work on chiral recognition using cyclodextrins is very famous and has been extensively reviewed.338−346 On the other hand, chiral recognition not only describes the interactions between different chiral molecules but also can represent the interactions between chiral supramolecular assemblies and chiral molecules. Herein, we focus on the chiral recognition of supramolecular assemblies.347−349 Thus, when molecules with contrary chirality interact with chiral supramolecular assemblies, the chiral recognition shows some new features, which are different from the case of chiral recognition between different chiral molecules. For example, when molecules with different chirality are mixed with chiral supramolecular gels, chiral recognition can be achieved from the change in appearance or rheological properties of the related supramolecular gels. More importantly, these chiral recognitions, detected as sol−gel transformation or color changes of the supramolecular gels, are visible to the naked eye.350,192 Although chiral recognition related to the host−guest chemistry of cyclodextrin has been widely investigated, chiral recognition based on highly symmetrical cucurbituril molecules is still intriguing. Thus, when achiral cucurbiturils form a very stable complex with chiral molecules, the corresponding assemblies are able to completely discriminate other enantiomers. This work has been reported by Inoue and coworkers.351 In this study, achiral cucurbiturils (CBs) were incorporated into (R)- or (S)-2-methylpiperazine, and the resulting complex showed significant enantiomeric discrimination of various chiral organic amines, such as (S)-2methylbutylamine (Figure 119). The chiral recognition could

Figure 119. Chiral recognition of (S)-2-methylbutylamine from achiral cucurbiturils (CBs) incorporated with chiral 2-methylpiperazine. Reprinted with permission from ref 351. Copyright 2006 American Chemical Society.

Although chiral recognitions based on cyclodextrins have been widely investigated, supramolecular polymers constructed from cyclodextrins also showed interesting recognition properties. For example, Yashima et al. synthesized polymers containing βcyclodextrin substituents and studied enantioselective gelation of these polymers in response to the chirality of a chiral amine. It has been found that (S)-1-phenylethylamine can help the polymers to form organogels but that (R)-1-phenylethylamine cannot.352 Shinkai et al. incorporated 4,4′-biphenyldicarboxylic acid into cyclodextrin (CD) as a bridging ligand, which further interacted with TbIII to form polyrotaxane-type metallosupramolecular polymers. Due to the chirality of cyclodextrin as well as the strong fluorescence of rare earth complexes, the recognition of small chiral molecules by this supramolecular polymer was achieved based on both fluorescence and circular dichroism spectral changes.353 One of the most important advantages of the chiral recognitions from different chiral soft matters can be the alteration of their macroscopic properties, which also can be 7366

DOI: 10.1021/cr500671p Chem. Rev. 2015, 115, 7304−7397

Chemical Reviews

Review

Figure 120. (A) Schematic illustration of the fabrication of a TbIII-based supramolecular polymer. (B) Fluorescence (ex. 270 nm) and CD spectra before and after adding D- or L-tartaric acid (final concentration; L, TbIII, and D/L-tartaric acid, 10 μm, α-CD, 300 μm). (dotted line) L + α-CD + TbIII, (dashed line) D-tartaric acid, and (solid line) L-tartaric acid. Reprinted with permission from ref 353. Copyright 2013 John Wiley & Sons.

Figure 122. Visual chiral recognition of binap through enantioselective metallogel collapsing. Reprinted with permission from ref 355. Copyright 2011 John Wiley & Sons.

Liu et al. synthesized novel L-glutamide-based amphiphilic gelators containing Schiff base moieties and long alkyl chains (oSLG, 8a and p-SLG, 8b). The self-assembly and gelation properties of these amphiphiles with different metal ions have been investigated in many organic solvents. In particular, adding metal ions to the organogels can significantly change the selfassembled nanostructures and the spectral characteristics of these systems. Thus, Cu2+ can transform the nanofiber gel into a chiral twist, while adding Mg2+ ions enhanced the fluorescence of the gels. Remarkably, organogels containing Mg2+ ions have very good chiral recognition ability. For example, when D- or L-tartaric acid was introduced into the system, the fluorescence quenching processes of these organogels can be totally different (Figure 123).158 Liu et al. designed and synthesized a series of amphiphilic molecules containing both glutamide moieties and long alkyl chains with different hydrophobic headgroups (Figure 5). The supramolecular assemblies of some of these amphiphiles can be used for the recognition of chiral molecules. Except for the organogels prepared from o-SLG and p-SLG, the assemblies based on quinolinol-functionalized L-glutamides (HQLG, 10) and metal ions were also found to have good chiral recognition capability.55 HQLG (10) can form complexes with different metal ions, such as Li+, Zn2+, and Al3+. Although these chiral complexes do not show a CD signal or chiral recognition properties in solution, they can form fluorescent metallogels with optical activity upon gelation in several organic solvents. The recognition of the small chiral organic molecules by these fluorescent metallogels can be detected by the changes in their CD spectra or fluorescence spectra. For example, metallogels formed by Zn2+/HQLG (10) complex showed a totally different fluorescent color when treated with (R,R)- or (S,S)-1,2-diaminocyclohexane. Therefore, using metallogels, chiral recognition can be achieved by the naked eye. It is believed that the self-assembled nanostructures play very important roles for chiral recognition (Figure 124). Porphyrins are very important molecular building blocks for the study of chiral supramolecular assembly. Ihara et al. synthesized an L-glutamide-functionalized zinc porphyrin (gTPP/Zn, 150) and investigated the enantioselective recognition of different amino acids by assemblies thereof (Figure 125A). gTPP/Zn (150) can form organogels in different organic solvents.

identified by the naked eye. Therefore, visual chiral recognitions can be realized. For example, in the case of supramolecular gels, the recognition of chiral gelators can be achieved from simply identifying whether the systems are a “gel”. For supramolecular gel systems, chiral recognition by the naked eye may be simply achieved from the sol−gel transformation. The work of Pu et al. is the first report of this phenomenon. They synthesized a Cu(II) terpyridine complex containing 1,1′-bi-2-naphthol (BINOL) substituents (148), which can form stable supramolecular gels in CHCl3 upon sonication. Interestingly, some chiral amino alcohols were found to change these gels into sols, while their enantiomers failed. Therefore, the chiral recognition of this organogel to some chiral amino alcohols can be achieved from the sol−gel transformation, which can be recognized by the naked eye. In this context, (S)phenylglycinol (0.10 equiv) can break the CHCl3 gel network, while (R)-phenylglycinol cannot. With chiral 1-amino-2propanol, the same enantioselective gel-collapsing process can also be observed (Figure 121).354

Figure 121. Chiral recognition from enantioselective gel collapsing, which formed from a Cu(II) terpyridine complex. Reprinted with permission from ref 354. Copyright 2010 American Chemical Society.

Tu et al. synthesized a gelator containing steroidal substituents and a platinum complex (149) and prepared metallogels from the assembly of these gelators. A visual chiral recognition can be realized from the sol−gel transformation of the metallogels. When (R)-binap was introduced into the system, the metallogels were destroyed. In contrast, (S)-binap cannot change the situation of metallogels (Figure 122).355 7367

DOI: 10.1021/cr500671p Chem. Rev. 2015, 115, 7304−7397

Chemical Reviews

Review

Figure 125. (A) Schematic image of enantioselective recognition through chirally ordered porphyrin assembly. (B) CD spectra of g-TPP/ Zn (150) (50 μM) with and without L- and D-His-OMe (50 μM) in cyclohexane at 20 °C. (C) Fluorescence spectra of the g-TPP/Zn (150) (50 μM) assembly with and without L- and D-His-OMe (50 μM) in cyclohexane at 20 °C. Reprinted with permission from ref 356. Copyright 2012 The Royal Society of Chemistry.

Figure 123. (A) Molecular structures and self-assembly of the amphiphilic Schiff bases with metal ions. (B) (a) Assembly mechanism of o-SLG (8a) and the chiral twist in the presence of Cu2+ ions. (b) oSLG (8a) formed a complex with Mg2+ ions and transferred the chirality to the whole assembly. When D-tartrate approached the Mg2+ ion, the Denantiomer was favored. (c) Cu2+ ions reacted with p-SLG (8b) and caused gelation. Reprinted with permission from ref 158. Copyright 2012 John Wiley & Sons.

results show that the coassembly of L-lysine dendrons with chiral amines can form supramolecular gel fibers, and the chirality of the amine could control the corresponding diastereomeric complexes. When both R and S amines were incorporated into the systems, the L-lysine dendrons could selectively coassemble with the R enantiomer, because the L-lysine dendron/R amine complexes formed the most stable gel. Moreover, when R amine enantiomers were added to the supramolecular gels formed by the L-lysine dendron and pure S amines, the diffusion of R amines and displacement of the original S amines from the “solid-like” fibers can be detected (Figure 126B).357 These results suggest that the two-component organogels are very sensitive to the molecular chirality of gelators and have great potential for chiral recognitions.

The CD spectra of the self-assembly of g-TPP/Zn in cyclohexane showed strong optical activity. Interestingly, L- and Denantiomers of many different α-amino acid derivatives can be differentiated by organogels of g-TPP/Zn. For example, when Lhistidine methyl ester (L-His-OMe) or D-histidine methyl ester (D-His-OMe) was mixed with g-TPP/Zn cyclohexane gels, very different CD and fluorescence spectra were obtained (Figure 125B and 125C).356 Smith and co-workers thoroughly investigated two-component organogels based on the coassembly of an L-lysine dendron (151) with different amines (Figure 126A). For these systems, when the chiral amines were used for the coassembly, very interesting chiral recognition phenomena can be detected. The

Figure 124. (A) Molecular structures of the HQLG ligand molecule (10) and its metal complexes. (B) Chiral recognition brought about by the metallogels. Reprinted with permission from ref 55. Copyright 2013 American Chemical Society. 7368

DOI: 10.1021/cr500671p Chem. Rev. 2015, 115, 7304−7397

Chemical Reviews

Review

substrates was determined using this porphyrin tweezer (Figure 127).368

Figure 127. Fluorinated porphyrin tweezer working as chiral sensor for the discrimination of the absolute configurations of amino alcohols. Reprinted with permission from ref 368. Copyright 2008 American Chemical Society.

Suzuki et al. synthesized secondary terephthalamide derivatives containing four aryl blades (153a-H), which can be used as the host molecule for chiral sensing. The conformation of this host can be changed from a nonpropeller anti form to a propellershaped syn form with the formation of a complex with some chiral molecules, such as p-xylylenediammonium derivatives. Moreover, depending on the chirality of the guest enantiomers coassembled with secondary terephthalamide derivatives, the complex can be biased to prefer a particular handedness, with the enhancement of CD signal. This secondary terephthalamide derivative can be used as chiral sensor for the discrimination of the very important neurotransmitter, (−)-phenylephrine (Figure 128).369 Some biomacromolecules can also be used for chiral recognition and sensing systems. As one of the most important biomacromolecules, DNA contains chiral information from the molecular to supramolecular level. Moreover, the stability of DNA can be a huge advantage for building different devices for chiral sensing. Qu et al. constructed electrochemical DNA sensors for the chiral sensing. In this system, DNA molecules modified with thiol groups were covalently bonded to a gold electrode. When small chiral molecules interacted with DNA, changes in the electrochemical characteristics of the gold electrode could be detected, thus demonstrating chiral sensing ability. It is worth mentioning that this system offers great advantages for distinguishing chiral metallosupramolecular complexes. For example, the authors reported a three-way junction based on an E-DNA sensor. In this study, the same palindromic DNA labeled with a redox-active methylene blue (MB) tag at the 5′-terminus was used as the support. Discrimination of chiral metallosupramolecular complexes with an enantioselective recognition ratio of about 3.5 was realized (Figure 129A and 129B).370 In another case, the authors reported a similar electrochemical DNA (E-DNA) chiral sensor based on the human telomeric G-quadruplex formation. An enantioselective recognition on zinc-finger-like chiral metallosupramolecular assemblies reaches ratios higher than 5 (Figure 129C).371 As previously described, supramolecular chiral assemblies have good capabilities for the recognition of low molecular weight chiral organic molecules. Worthy of note are the chiral supramolecular assemblies that can recognize very small amounts of chiral molecules. In this context, we will show some representative examples of chiral sensing based on supramolecular assemblies.

Figure 126. (A) Chiral gelation system of an L-lysine dendron (G2-Lys, 151) and chiral amines (C6R/S). (B) Schematic of thermodynamically controlled gel evolution upon addition of C6R to a gel made from G2Lys (151) and C6S. Reprinted with permission from ref 357. Copyright 2014 American Chemical Society.

In principle, if the chiral recognition can be triggered by a very small amount of chiral organic molecules, these assemblies with chiral recognition ability can be used as a chiral sensor. Compared with the recognition of chiral molecules, chiral sensors based on supramolecular assemblies can be more difficult to prepare. On the other hand, some wide-sense chiral sensors are available based on organic molecules, polymers, and supramolecular assemblies.192,358,359 Herein, we discuss some typical examples of chiral sensors that have been recently reported. Very simple but efficient chiral sensors can be prepared from the biaryls, which have been developed mainly by Wolf et al.360,361 For example, naphthalene derivatives containing salicylaldehyde units and pyridyl N-oxide fluorophores have been synthesized. This is the result of the salicylaldehyde unit which can form a complex with amino alcohols with different chirality and subsequently change the conformation of the molecules. The chiral sensing can be determined from the changes in the CD and fluorescence spectra. By using this system, the absolute configuration of many amino alcohols can be analyzed.42 Covalently connected porphyrin dimers, which were named porphyrin tweezer systems, have been thoroughly studied by Nakanishi, Berova, and co-workers. These porphyrin tweezer systems can be used to determine the stereochemistry of many small chiral organic molecules.362−367 Indeed, some of the porphyrin tweezers can be very good chiral sensors with determination of the absolute chirality of the guest molecules. For example, Borhan et al. designed and synthesized an electron-deficient fluorinated porphyrin tweezer (152) and demonstrated that it is a good chiral sensor for the discrimination of many different small organic molecules containing two chiral centers. Depending on the chirality of the small organic substrates, the changes in the CD spectrum of the supramolecular assemblies formed by the porphyrin tweezers and chiral guests can be detected. In this case, the absolute stereochemical configuration of a variety of erythro and threo 7369

DOI: 10.1021/cr500671p Chem. Rev. 2015, 115, 7304−7397

Chemical Reviews

Review

Figure 128. (A) Molecular structures of the secondary terephthalamide derivatives containing four aryl blades (153) and their different conformations. (B) Schematic image showing the conformational changes from a nonpropeller anti form to a propeller-shaped syn form upon forming complexes with chiral guests. Reprinted with permission from ref 369. Copyright 2009 American Chemical Society.

Figure 129. (A) Schematic illustration of a three-way junction based on E-DNA for distinguishing chiral metallo-supramolecular complexes. (B) Changes of current from the E-DNA sensor showing discrimination of chiral metallosupramolecular complexes. (C) Schematic representation of the human telomeric DNA-based electrochemical DNA (E-DNA) sensor. Reprinted with permission from refs 370 and 371. Copyright 2012 The Royal Society of Chemistry.

Figure 130. Self-assembly of zinc porphyrin dimers can lead to box-shaped tetramers, which shows chiroptical sensing for limonene. Reprinted with permission from ref 372. Copyright 2007 John Wiley & Sons.

substituents (154). This zinc porphyrin dimer can self-assemble into box-shaped tetramers. In a solution of asymmetric hydrocarbons, such as limonene, the self-assembly of the zinc

Porphyrins are very important building blocks for the study of supramolecular chirality. Tsuda and Aida synthesized a zinc porphyrin dimer containing a rigid linker and pyridine 7370

DOI: 10.1021/cr500671p Chem. Rev. 2015, 115, 7304−7397

Chemical Reviews

Review

Figure 131. (A) Four-component reversible covalent assembly for secondary alcohol binding. OTf is trifluoromethanesulfonate (triflate). (B) Exploration of four-component assembly for chirality sensing and ee determination. CD spectra of 1-phenylethanol-induced assembly with different ee’s of the alcohol (from top to bottom −100%, −80%, −60%, −40%, −20%, 0%, 20%, 40%, 60%, 80%, 100%). Reprinted with permission from ref 373. Copyright 2011 Nature Publishing Group.

porphyrin dimer can lead to a homochiral box-shaped tetrameric assembly. This self-assembled porphyrin box is enantiomerically enriched and optically active, showing that chiroptical sensing can be realized (Figure 130). From the CD spectra of the porphyrin box, the absolute configuration of limonene can be determined. Interestingly, in the case of very small enantiomeric enrichment of limonene, extremely large molecular ellipticities of the porphyrin boxes were detected.372 Chiral sensing based on multicomponent assemblies containing reversible covalent bonding are also worth mentioning. Anslyn et al. constructed a four-component reversible assembly containing carbonyl activation and hemiaminal ether stabilization (155) (Figure 131A). In this system, reversible binding of the monoalcohol has been achieved. Moreover, the tetradentate ligand of the assembly renders close incorporation of secondary alcohols. By binding and exchange of chiral alcohols, chiral sensing can be demonstrated by CD spectral measurements (Figure 131B). This chiral sensor can be used for determination of the enantiomeric excess of mixed chiral alcohols with different ee values (Figure 131B).373 A chiral sensor with the ability to differentiate many different chiral molecules, such as amino acids, peptides, proteins, and even some aromatic drugs, has been developed by Biedermann and Nau. These systems are based on ternary complexes formed between the macrocyclic host cucurbit[8]uril, dicationic dyes, and chiral aromatic analytes (Figure 132A). Although both cucurbit[8]uril and dicationic dyes are achiral, the chirality of the aromatic analytes with low micromolar concentrations in water can be detected by the changes in the CD spectra. Remarkably, by using these chiral sensors, peptide sequences can also be recognized. Most interestingly, since the chiral sensors are constructed via noncovalent interactions with good reversibility, real-time monitoring of the chirality of analytes can be achieved. Therefore, for certain enzyme-catalyzed chemical reactions, the rate of reactions, the yield of products, and the ee values of the products can be detected in real time by measuring the CD spectra (Figure 132B).374 Although we would like to focus on supramolecular chirality within self-assembled systems in this review, we still need to cover some of the chiral metal nanomaterials in this portion to fully address chiral sensing. Because of the free electrons on the surface, the plasmonic absorption or CD spectra from the chiral metal nanomaterials can be very sensitive for detecting the interactions with other chiral molecules,375−380 and chiral metal nanomaterials can be constructed as ultrasensitive chiral sensors.

Figure 132. (A) Schematic illustration of the chiral sensing based on the ternary complexes between the macrocyclic host cucurbit[8]uril, dicationic dyes, and chiral aromatic analytes. (B) Examples of reaction monitoring. Reprinted with permission from ref 374. Copyright 2014 John Wiley & Sons.

For example, Kadodwala et al. built an ultrasensitive chiral sensor based on chiral metamaterials, which had the potential to discriminate between large biomolecules with similar levels of sensitivity and subtle structural differences at pictogram quantities. In this study, the optical excitation of plasmonic planar chiral metamaterials can generate superchiral electromagnetic fields, which are highly sensitive for the detection of chiral peptide nanostructures. The sensitivity of this chiral sensor was found to be up to 106 times greater than that of the optical polarimetry measurements. The largest differences were observed for proteins with high β-sheet content. The system 7371

DOI: 10.1021/cr500671p Chem. Rev. 2015, 115, 7304−7397

Chemical Reviews

Review

Figure 133. Changes induced in the chiral plasmonic resonances of the planar chiral metamaterials (PCM) are readily detected using CD spectroscopy. (A) CD spectra collected from PCMs immersed in distilled water. (B) Influence of the adsorbed proteins hemoglobin, β-lactoglobulin, and thermally denatured β-lactoglobulin on the CD spectra of the PCMs. (C) Hemoglobin (top) and β-lactoglobulin (bottom) (α-helix, cyan cylinder; β-sheet, ribbons), shown adopting a well-defined arbitrary structure with respect to a surface. The figure illustrates the more anisotropic nature of adsorbed βlactoglobulin. Reprinted with permission from ref 381. Copyright 2010 Nature Publishing Group.

has great potential for detecting amyloid diseases and certain types of viruses (Figure 133).381 6.2. Supramolecular Chiroptical Switches

In the field of supramolecular chemistry and nanotechnology, constructing assemblies that act like a “switch” is one of the most interesting topics. In general, any assembly in which specific properties can be reversibly changed with an external stimulus can be regarded as a supramolecular switch. A supramolecular chiroptical switch is based on reversible changes of supramolecular chirality, which is seen as externally stimulated optical activity.382−384 The changes of supramolecular chirality can be achiral to chiral, either reversible or reversible from left-handed chirality to right-handed chirality. Various supramolecular assemblies, such as liquid crystals,385 host−guest complexes,386 LB films,387−389 and supramolecular gels,51,56 have been developed as supramolecular chiral switches. In addition, some small organic molecules or polymers can also be used as chiroptical switches based on photoirradiation or interaction with other small molecules.391−396 A supramolecular chiroptical switch based on an amorphous azobenzene polymer (156) has been constructed by Kim et al. When the thin films of an achiral epoxy-based polymer containing photoresponsive azobenzene groups were irradiated by elliptically polarized light (EPL), supramolecular chirality was introduced into the system, which was confirmed by the CD spectral measurements. The helical arrangement of the azobenzenes plays a very important role in the photoinduced supramolecular chirality. When the irradiation at 488 nm was changed from right-handed elliptical polarization to left-handed elliptical polarization, the handedness of the chiral supramolecular assembly also changed reversibly. This supramolecular chiroptical switch was able to operate several times before fatigue resistance occurred (Figure 134).391 Polythiophene is a very important conductive polymer. Yashima et al. constructed the first reversible supramolecular chirality switch based on chiral polythiophene aggregates. When copper(II) trifluoromethanesulfonate [Cu(OTf)2] was added to chiral aggregates of chiral regioregular polythiophene in a chloroform−acetonitrile mixture, the CD signal of the assemblies disappeared due to the oxidative doping of the polymer main chain. However, when amines, such as triethylenetetramine (TETA), were added to the system to undope the polymer, the CD signals reappeared. Thus, a supramolecular chiral switch can be prepared from the assembly of chiral polythiophenes with the

Figure 134. (A) Chemical structure of an achiral epoxy-based polymer containing photoresponsive azobenzene groups (156) and the helical arrangement of the azobenzenes with different handedness. (B) (a) Chiroptical switching of the CD spectra by alternating irradiation with rand l-EPL. (b) Intensity of the CD signal at 410, 510, and 700 nm. Reprinted with permission from ref 391. Copyright 2006 John Wiley & Sons.

addition or removal of an electron from the corresponding polymer main chain.397 Chiral LS films constructed from the air/water interfacial assembly of achiral molecular building blocks can be used as supramolecular chiral switches. In this context, the handedness of the supramolecular chirality of assemblies constructed by achiral molecules can be changed by an external stimulus. This flexibility 7372

DOI: 10.1021/cr500671p Chem. Rev. 2015, 115, 7304−7397

Chemical Reviews

Review

is a typical characteristic of assemblies based on noncovalent interactions, and supramolecular chiral switches have been achieved from the assembly of amphiphilic molecules on a water surface. For example, achiral 5-(octadecyloxy)-2-(2thiazolylazo)phenol (TARC18, 157) can form chiral LS films via an air/water interfacial assembly. Supramolecular chiral switches based on the LS films of achiral TARC18 were fabricated by alternately exposing the film to HCl gas and to air (Figure 135).387

Figure 136. (A) Molecular structure of azobenzene-substituted diacetylene (NADA, 158). (B) Schematic illustrations of (a) enantioselective polymerization with CPUL irradiation and (b) chirality modulation for NADA LB films with CPL treatment. Reprinted with permission from ref 388. Copyright 2009 The Royal Society of Chemistry. Figure 135. (A) Molecular structure of TARC18 (157). (B) Schematic illustration of the possible helical stacking of the TARC18 (157) molecules in the LS films. M and P chiralities of the films were formed by chance. Reprinted with permission from ref 387. Copyright 2006 John Wiley & Sons.

their well-known work on light-driven molecular motors.399−401 In this context, the modified light-driven molecular motor was covalently connected to the terminus of poly(n-hexyl isocyanate) (PHIC), and the chiral information from the headgroups of the polymer was found to be expressed at both the supramolecular and the macromolecular levels. Upon irradiation with two different wavelengths of light, a chiroptical switch has been demonstrated, as evidenced by the CD spectral measurements and images obtained from an optical microscope equipped with crossed polarizers (Figure 137).385

Chiral thin films that were constructed from achiral phthalocyanine derivatives via air/water interfacial assembly can also be used as supramolecular chiral switches.382 In particular, supramolecular chiroptical switches based on the assembly of achiral phthalocyanine derivatives were found to be very stable. Certainly, the π−π interactions between phthalocyanine rings can increase the stability of the assembly. Most importantly, the polymerization of achiral phthalocyanine derivatives can produce chiral assemblies based on covalent bonds. When LS films containing polymerized chiral assemblies of achiral phthalocyanine derivatives were alternately exposed to HCl and NH3, a reversible change in the CD spectra was detected. This process can be repeated many times without any decrease in CD signal intensity. Zou et al. constructed LB films of azobenzene-substituted diacetylene (NADA, 158). Although NADA is achiral, the NADA LB films show supramolecular chirality. Moreover, the assemblies within the LB films can be polymerized with photoirradiation. When left- and right-handed circularly polarized ultraviolet light (CPUL) was applied, polymerized NADA (PNADA) LB films with different chirality were obtained, as confirmed by the corresponding CD spectra from both azobenzene chromophores and polydiacetylene (PDA) chains. Interestingly, these polymerized LB films NADA (PDA LB films) can be used as chiroptical switches after irradiation using left- and right-handed circularly polarized lasers (CPL, 442 nm), due to alternation of the stereoregular packing of azobenzene chromophores.388 The self-assembly of banana-shaped achiral molecules was found to lead to chiral liquid crystals, and chiroptical switches based on liquid crystals containing only achiral molecular building blocks were also achieved. Tschierske et al. synthesized bent-core mesogens carrying branched oligosiloxane units. A chiroptical switch based on the phase transition was obtained by applying an electric field or changing the temperature.398 An elegant light-driven supramolecular chiroptical switch was developed by Feringa and co-workers. This work was based on

Figure 137. Schematic representation of the full photocontrol of the magnitude and sign of the supramolecular helical pitch of a cholesteric LC phase generated by a polyisocyanate with a single chiroptical molecular switch covalently linked to the polymer’s terminus. Reprinted with permission from ref 385. Copyright 2008 American Chemical Society.

Chiral polymers containing azobenzene groups were also found to form liquid crystals. Chiral switches based on these chiral liquid-crystalline polymers were investigated by Angiolini and co-workers. It was found that the chiral polymers containing azobenzene groups and L-lactic acid formed a smectic A1/2 (fully interdigitated) liquid-crystalline phase. The resulting chiroptical switching of the system was achieved by irradiation with circularly polarized light (CPL) with different handedness.402 Chiroptical switches based on some soft matters, such as supramolecular gels or supramolecular assemblies in solution, have attracted increased attention recently. Cucurbituril is a very important building block for preparing chiral supramolecular assemblies. Supramolecular chiroptical switches based on the coassembly of chiral binaphthalene−bipyridinium guests together with cucurbituril hosts have been developed by Venturi and 7373

DOI: 10.1021/cr500671p Chem. Rev. 2015, 115, 7304−7397

Chemical Reviews

Review

Tian et al.403 In this study, three molecular tweezers containing 4,4′-bipyridinium (BPY2+) and (R)-2,2′-dioxy-1,1′-binaphthyl (BIN) units with different length alkyl chains were designed and synthesized (159). In aqueous solution, these chiral binaphthalene−bipyridinium guests formed complexes with cucurbituril hosts. In this system, chiroptical switching was achieved from the reversible changes of the helicity of the BIN units, which was triggered by reduction of the BPY2+ units. In addition, the alkyl linkers also play an important role in association to cucurbiturils. Thus, changing the length of the alkyl chains modulated the properties of the chiral switches, such as the molar ratio of the complex and the dihedral angle of BIN (Figure 138). Figure 139. Schematic illustration of the self-assembled azobenzenecontaining lipid showing multiresponsibility for chiral switching. Reprinted with permission from ref 56. Copyright 2011 American Chemical Society.

nanotubes, this hierarchical assembly cannot be used as a chiroptical switch. Only when Azo was coassembled with the bolaamphiphile at the molecular level could the corresponding assemblies show reversible changes both in the UV−vis and in the CD spectra upon alternative UV−vis irradiation of Azo (Figure 140).404 Chiral supramolecular nanotubes assembled by L- or Dglutamic-acid-based bolaamphiphiles (HDGA, 15b) in water can also be used as templates to produce silica nanotubes. Most interestingly, only the inner walls of the formed silica nanotubes were found to have supramolecular chirality. When the photoactive azobenzene moieties were loaded onto the inner chiral silica nanotubes, chiroptical switches based on inorganic nanomaterials resulted (Figure 141).27 6.3. Supramolecular Chiral Catalysis

In synthetic chemistry, construction of chiral molecules has been found to be extremely important. Therefore, developing different chiral homogeneous catalysts and heterogeneous catalysts has become a significant research goal.405−411 For example, coordination complexes containing metal ions and chiral ligands have been widely investigated as highly efficient chiral catalysts. Many outstanding reviews have been published examining this issue.410−416 In particular, as a class of very important coordination complexes with crystalline structures, metal− organic frameworks (MOFs), which have infinite network structures built with multitopic organic ligands and metal ions, have been thoroughly studied as potential asymmetric catalysts. This topic has also been extensively reviewed recently.417 Similarly, the catalytic properties of cyclodextrin derivatives have also attracted increased attention recently.418,419 As the chiral host, cyclodextrin derivatives can provide a suitable chiral environment, and reactions at the guest molecules can provide reaction products with chirality. Moreover, the molecular structures of cyclodextrin derivatives can be further modified to obtain more efficient chiral catalysts.420 There are also many good review articles concerning chiral catalysis based on cyclodextrin derivatives. For example, Inoue and co-workers summarized supramolecular photochirogenesis based on cyclodextrin derivatives.421,422 In a general sense, “supramolecular chiral catalysis” is presently a major topic of research interest, but we cannot address every aspect of this field of research. In this review, we concentrate on the catalytic properties of some chiral supramolecular assemblies. Although the catalytic properties of metal complexes have been

Figure 138. (A) Structure of chiral binaphthalene−bipyridinium guests (159). (B) Pictorial representation of the conformational rearrangements of 159 in response to two-electron reduction and/or complexation with either CB[8] or CB[7]. Reprinted with permission from ref 403. Copyright 2012 John Wiley & Sons.

The very good stimulus-responsive properties of selfassembled systems based on noncovalent interactions can help these systems form chiral switches with unique performance. For example, Liu et al. synthesized a glutamic-acid-based lipid containing an azobenzene headgroup (azo-LG2C18, 5), which can form organogels with supramolecular chirality in different organic solvents. Remarkably, the resulting organogels can be used as chiroptical switches with multiresponsibility. Thus, the supramolecular chirality can be changed reversibly by photoirradiation, temperature variation, or solvent polarity (Figure 139).56 A supramolecular chiroptical switch based on multicomponent self-assembled soft matters has also been developed by Liu group. The self-assembly of chiral glutamic-acid-based bolaamphiphiles (HDGA, 15b) can lead to hydrogels and chiral supramolecular nanotubes. For the construction of multicomponent soft matters, the azobenzene derivative 4(phenylazo)benzoic acid sodium salt (Azo) can be coassembled with either HDGA molecules (15b) or chiral supramolecular nanotubes. Although very strong supramolecular chirality can be detected from the coassembly of Azo with the preformed chiral 7374

DOI: 10.1021/cr500671p Chem. Rev. 2015, 115, 7304−7397

Chemical Reviews

Review

Figure 140. (A) Structures of the hydrogelators HDGA (15b) and Azo. (B) Gels formed by L-HDGA in water in the absence and presence of Azo. (C) Illustration of the coassembly of Azo with HDGA. Reprinted with permission from ref 404. Copyright 2011 The Royal Society of Chemistry.

induction and reusability in the catalysis of asymmetric hydrogenation of dehydro-α-amino acid and enamide derivatives (Figure 142).423 Furthermore, Ding et al. synthesized a heteroditopic ligand containing a 2,2′:6′,2″-terpyridine (tpy) unit and Feringa’s MonoPhos (160). The selective coordination of this ligand with FeII and RhI ions can produce chiral supramolecular polymers, which can be used as chiral bimetallic self-supported catalysts. In the hydrogenation of α-dehydroamino acid, enamide, and itaconic acid derivatives, these reusable heterogeneous asymmetric catalysts can lead to very high reaction rates with excellent enantioselectivity (90−97% ee) (Figure 143).424 Among various supramolecular chiral catalysis from selfassembly, ion pair catalysts, which have been developed in recent years, are very important systems with many distinctive characteristics. Ooi et al. developed supramolecular assemblies containing ion pairs through intermolecular hydrogen bonding. This system was formed by the coassembly of a chiral tetraaminophosphonium cation, two phenols, and a phenoxide anion and was found to have chiral catalytic activity (161). In solution, this ion pair complex promotes a highly stereoselective conjugate addition of acyl anion equivalents to α,β-unsaturated ester surrogates with a broad substrate scope (Figure 144).425 Ishihara et al. also produced self-assembled chiral catalysts without metal ions. These supramolecular catalysts are based on in situ coassembly from chiral diols, arylboronic acids, and tris(pentafluorophenyl)borane (162). In the Diels−Alder reactions of cyclopentadiene with different acroleins, these

Figure 141. Creating chirality in the inner walls of silica nanotubes through a hydrogel template, and the chiroptical switching of these nanotubes. Reprinted with permission from ref 27. Copyright 2010 The Royal Society of Chemistry.

well investigated, supramolecular polymers based on coordination interactions have very special characteristics for supramolecular chiral catalysis. For example, Ding et al. constructed polymeric supramolecular chiral catalysts based on self-assembly. The authors synthesized an organic ligand containing ureido-4[1H]ureidopyrimidone (UP) and Feringa’s MonoPhos motifs. By mixing the organic ligand with [Rh(cod)2]BF4, supramolecular polymers can be produced by orthogonal self-assembly via hydrogen-bonding and ligand-to-metal coordination interactions. This supramolecular polymer shows excellent asymmetric

Figure 142. (A) Polymeric supramolecular chiral catalyst based on self-assembly. (B) Asymmetric hydrogenation of dehydro-α-amino acid and enamide derivatives. Reprinted with permission from ref 423. Copyright 2006 John Wiley & Sons. 7375

DOI: 10.1021/cr500671p Chem. Rev. 2015, 115, 7304−7397

Chemical Reviews

Review

Figure 143. (A) Molecular structure of a ligand containing the 2,2′:6′,2″-terpyridine (tpy) unit and Feringa’s MonoPhos (160), and schematic illustration of supramolecular catalysts through orthogonal coordination of two different metal ions with a single ditopic ligand. (B) Asymmetric hydrogenation of dehydroamino acid, enamide, and itaconic acid derivatives using a catalyst with high reaction rate and excellent enantioselectivity. Reprinted with permission from ref 424. Copyright 2010 John Wiley & Sons.

Supramolecular catalysts based on the coassembly of chiral amines and poly(alkene glycol)s have been reported by Xu et al. These systems were found to be highly efficient in the asymmetric catalysis of the unusual Diels−Alder reaction between cyclohexenones and nitrodienes, nitroenynes, or nitroolefins, providing excellent chemo-, regio-, and enantioselectivities (Figure 146).427

Figure 146. Coassembly of chiral amines and poly(alkene glycol)s showing highly efficient asymmetric catalysis of Diels−Alder reactions. Reprinted with permission from ref 427. Copyright 2011 John Wiley & Sons.

Although many chiral supramolecular catalysts have been developed via self-assembly, the chiral catalysis characteristics of systems are still largely dependent on chirality at the molecular level. By contrast, even if many chiral nanostructures have been constructed via self-assembly, the relationship between chirality at the nanoscale and chiral catalysis at the molecular level is still rarely discussed. As mentioned previously, the self-assembly of chiral glutamicacid-based bolaamphiphiles (HDGA, 15b) led to hydrogels and chiral supramolecular nanotubes. When Cu2+ ions were added to the system, a monolayer nanotube was transformed into a

Figure 144. (A) Structures of chiral tetraaminophosphonium cations. (B) Oak Ridge thermal ellipsoid plot diagram of 161a·(OPh)3H2. (C) Scope of α,β-unsaturated acylbenzotriazole. Reprinted with permission from ref 425. Copyright 2009 The American Association for the Advancement of Science.

catalysts have very good endo/exo selectivities and high enantioselectivities (Figure 145).426

Figure 145. (A) Structures of chiral supramolecular catalyst (162). (B) Enantioselective Diels−Alder reactions with anomalous endo/exo selectivities using chiral supramolecular catalysts. Reprinted with permission from ref 426. Copyright 2011 John Wiley & Sons. 7376

DOI: 10.1021/cr500671p Chem. Rev. 2015, 115, 7304−7397

Chemical Reviews

Review

multilayer nanotube with a tubular wall thickness of about 10 nm. Interestingly, the resulting Cu2+-containing supramolecular nanotubes were useful as an asymmetric catalyst for the Diels− Alder reaction between cyclopentadiene and aza-chalcone, which accelerates the reaction rate and enhances enantiomeric selectivity. Thus, asymmetric catalysis of the molecular reaction can be achieved by chiral nanostructures. It was suggested that through the Cu2+-mediated nanotube formation the substrate molecules could be anchored on the nanotube surfaces to produce a stereochemically favored alignment (Figure 147). Therefore, when adducts reacted with the substrate, both the enantiomeric selectivity and the reaction rate were found to increase.428

Figure 148. Self-assembly of vesicles regulated by compressed CO2 and the proposed transition-state model for the direct asymmetric aldol reaction. Reprinted with permission from ref 429. Copyright 2013 John Wiley & Sons.

Yashima et al. synthesized a chiral polymer containing riboflavin units as the main chain (165). Within the polymers, the 5-ethylriboflavinium cations can be reversibly transformed into 4a-hydroxyriboflavins upon hydroxylation/dehydroxylation, which renders significant changes in the absorption and circular dichroism (CD) spectra of the polymers. It is believed that the face-to-face stacking of the intermolecular riboflavinium units within the polymer produced twisted helical nanostructures with supramolecular chirality. This optically active polymer containing 5-ethylriboflavinium cations was found to efficiently catalyze the asymmetric organocatalytic oxidation of sulfides with hydrogen peroxide, yielding optically active sulfoxides with up to 60% ee (Figure 150).431 DNA is a very important genetic material of living organisms. However, it can also be regarded as a very useful chiral functional polymer for different applications. As a polymer, DNA has many different molecular chiral centers and charged substituents, which can increase its solubility in water. Moreover, the folding of DNA can produce very ordered nanostructures via hydrogen bonding, which can also be modulated by changing the sequences in the DNA. It should be noted that DNA is relatively stable in comparison to other biomacromolecules, such as RNA. Therefore, developing DNA-based supramolecular chiral catalysts has recently attracted interest. The first DNA-based asymmetric catalyst containing copper ions was reported by Roelfes and Feringa. In this study, the authors synthesized an achiral ligand containing a DNAintercalating moiety (9-aminoacridine), alkyl chain spacer, and metal-binding group (166). A DNA-based asymmetric catalyst was fabricated from the coassembly of a copper(II)-enclosing ligand and DNA. In the Diels−Alder reaction, this chiral catalyst could transfer the chirality of the DNA into the products, with the an ee value up to 90% (Figure 151A).432 For many other organic chemical reactions, a DNA-based chiral catalyst containing copper(II) was also found to be very useful. For example, Roelfes et al. developed a DNA-based asymmetric catalyst containing copper(II) and an achiral ligand for catalyzing the Michael reaction in water to achieve high enantioselectivity. These reactions can be performed on a relatively large scale, allowing recycling of the supramolecular chiral catalyst. For this system, many simple achiral ligands, such

Figure 147. Illustration of the assembly mechanism of the Cu2+−LHDGA nanotubular structure and its asymmetric catalysis of the Diels− Alder reaction of aza-chalcone with cyclopentadiene. Reprinted with permission from ref 428. Copyright 2011 American Chemical Society.

Besides catalytic nanotubes, chiral catalysis based on supramolecular nanostructures has also been observed using vesicles. Liu et al. synthesized amphiphilic molecules containing a proline headgroup (PTC12, 163). The self-assembly of PTC12 in water under compressed CO2 can produce vesicles. These assemblies were found to catalyze the asymmetric aldol reaction with high enantiomeric selectivity without any additives. Importantly, the size of the PTC12 assemblies and subsequently catalyst activity and stereoselectivity can be dynamically modulated by changing the status of the compressed CO2. Moreover, because CO2 can be easily removed from the system, it is very convenient for the separation and purification of products, as well as the reuse of the chiral supramolecular catalysts (Figure 148).429 In the development of supramolecular chiral catalysis, chiral covalent polymers, including some biomacromolecules, such as DNA and polypeptides, can be used as building blocks. The catalytic capability of these polymers may originate from the molecular chiral centers within these polymers but may also result from their folding characteristics and hierarchical nanostructures. Meijer and co-workers synthesized water-soluble segmented terpolymers containing PEG and chiral benzene-1,3,5-tricarboxamide side chains as well as a ruthenium complex (164). Due to the chiral self-assembly of the benzene-1,3,5-tricarboxamide side chains, the folding of these polymers can produce a helical structure in the apolar core around a ruthenium-based catalyst. This catalyst, resulting from the folding of polymers, was found to catalyze the transfer hydrogenation of ketones (Figure 149).430 7377

DOI: 10.1021/cr500671p Chem. Rev. 2015, 115, 7304−7397

Chemical Reviews

Review

Figure 149. (A) Water-soluble segmented terpolymer containing PEG and chiral benzene-1,3,5-tricarboxamide side chains as well as a ruthenium complex (164). (B) Supramolecular single-chain folding of polymers in water affording a compartmentalized catalyst for the transfer hydrogenation of ketones. Reprinted with permission from ref 430. Copyright 2011 American Chemical Society.

Figure 150. Optically active polymers consisting of riboflavin units catalyze the asymmetric organocatalytic oxidation of sulfides. Reprinted with permission from ref 431. Copyright 2012 American Chemical Society.

sulfones in water, high enantioselectivities with ee values of up to 84% were achieved.435 Most interestingly, use of DNA-based chiral catalysts containing copper(II) for chemical reactions in biological systems, such as enantioselective addition of water to olefins in an aqueous environment, resulted in good efficiency under the laboratory conditions. For the enantioselective hydration of enones, the chiral β-hydroxy ketone product can be obtained with an ee up to 82% upon catalysis with DNA-based assemblies (Figure 151D). Moreover, the reaction was also found to be diastereospecific, with the formation of only the syn hydration product.436 The folding of DNA can produce different nanostructures, which can also be modulated by changing the DNA sequences or other conditions for assembly. Besides double-helix DNA, telomeric G-quadruplex DNA was also studied to construct DNA-based chiral catalysts. Moses et al. constructed supramolecular chiral catalysts based on the assembly of telomeric Gquadruplex DNA, achiral ligands, and copper(II). These catalytic systems were found to catalyze Diels−Alder reactions successfully with modest enantioselectivities.437 Another class of G-quadruplex-DNA-based chiral catalysts was developed by Li and co-workers. These supramolecular chiral catalysts were constructed by self-assembly of human telomeric G4DNA and different metal ions. In this case, additional achiral ligands were not needed for building the catalysts. In an asymmetric Diels−Alder reaction, the complex of human telomeric G4DNA and Cu2+ ions provided a significant enhancement in the reaction rate with good enantioselectivity

as dipyridine, can be used for coassembly with DNA and copper ions (Figure 151B). The reactants in this study included α,βunsaturated 2-acylimidazoles working as the Michael acceptors and nitromethane and dimethyl malonate as the nucleophiles. Upon chiral catalysis by the DNA-based assemblies containing copper(II) and achiral ligand, the enantioselectivities of the Michael reaction were found to be up to 99% ee.433 Furthermore, Feringa and Roelfes also studied asymmetric Friedel−Crafts alkylation with olefins in water catalyzed by a DNA-based chiral catalyst containing copper(II). In this system, 4,4′-dimethyl-2,2′-bipyridine (dmbpy) was used as the achiral ligand for the complex with copper(II) and coassembly with DNA. For the asymmetric Friedel−Crafts reaction of α,βunsaturated 2-acylimidazoles with heteroaromatic π nucleophiles, good yields and high enantioselectivities were obtained using a very small of amount of DNA-based chiral catalysts (Figure 151C). In this study, the catalytic efficiency of both double-stranded DNA and single-stranded DNA with different sequences was investigated. The results showed that only the chiral catalysts assembled by the double-stranded DNA can introduce high enantioselectivities by catalyzing the asymmetric Friedel−Crafts alkylation. In addition, the highest enantioselectivities (up to 93%) were obtained by the supramolecular catalysts assembled using d(TCAGGGCCCTGA)2 DNA.434 A DNA-based chiral catalyst containing copper(II) and achiral ligand was also studied in the reaction of asymmetric intramolecular cyclopropanation (Figures 151−155). For the asymmetric intramolecular cyclopropanation of α-diazo-β-keto 7378

DOI: 10.1021/cr500671p Chem. Rev. 2015, 115, 7304−7397

Chemical Reviews

Review

Figure 151. (A) Asymmetric Diels−Alder reaction of cyclopentadiene with aza-chalcone, catalyzed by copper complexes of the ligand in the presence of DNA. (B) Asymmetric Michael addition reaction catalyzed by complexes formed between copper(II) ions and achiral ligands in the presence of DNA. (C) Cu−dmbpy/st-DNA-catalyzed Friedel−Crafts alkylation. Reprinted with permission from refs 432, 433, and434. Copyright 2005, 2007, and 2009 John Wiley & Sons. (D) DNA-based catalyst and general reaction scheme of the catalytic enantioselective hydration of a variety of α,β-unsaturated 2acyl-(1-alkyl)imidazole substrates, and overview of ligands used in this study. Reprinted with permission from ref 436. Copyright 2010 Nature Publishing Group. (E) Intramolecular cyclopropanation of α-diazo-β-keto sulfones in water using a DNA-based catalyst. Reprinted with permission from ref 435. Copyright 2013 The Royal Society of Chemistry.

Figure 152. Enantioselective Diels−Alder reactions with G-quadruplex-DNA-based catalysts; the absolute configuration of the products can be reversed when the conformation of the G4DNA is switched from antiparallel to parallel. Reprinted with permission from ref 438. Copyright 2012 John Wiley & Sons.

(74% ee). In addition, the rate and enantioselectivity of the reaction can be modulated by changing the DNA sequence and metal ions used to form the complex. Interestingly, the absolute configuration of the products can be controlled by the assembly of chiral catalysts (Figure 152). Thus, when the conformation of the G4DNA was switched from antiparallel to parallel, the absolute configuration of the products obtained from Diels− Alder reactions could be reversed.438

For the DNA-based chiral catalysis, the relationship between the handedness of the DNA helix and the molecular chirality of products was investigated by Smietana and Arseniyadis et al. They constructed different DNA-based supramolecular chiral catalysts from the assembly of both L-DNA and D-DNA. The LDNA, which contains deoxyribose with an L-conformation, can self-assemble into left-helical nanostructures, while the folding of normal DNA only produces right-helical nanostructures. Therefore, the L-DNA-based and D-DNA-based supramolecular chiral 7379

DOI: 10.1021/cr500671p Chem. Rev. 2015, 115, 7304−7397

Chemical Reviews

Review

Figure 153. Tuning the absolute configuration in DNA-based asymmetric catalysis. Reprinted with permission from ref 439. Copyright 2013 John Wiley & Sons.

covalent interactions between amino acids during the folding of polypeptides, which produce more flexible nanostructures. For supramolecular chiral catalysis under laboratory conditions, such complicated and often unstable nanostructures are not easy to handle. Therefore, even though nature has used enzymes for catalyzing many very subtle chemical reactions for billions of years in catalyzing chemical reactions under artificial conditions, DNA is still a better candidate. Nevertheless, polypeptide-based chiral catalysts have also been constructed recently. For example, Roelfes et al. modulated natural bovine pancreatic polypeptides with nonproteinogenic amino acids for binding Cu2+ ions. The resulting metalloenzymes catalyze Diels−Alder and Michael addition reactions in water with high enantioselectivities (Figure 155).441 Herrmann et al. developed chiral catalysts based on natural polypeptides without many chemical modifications. These polypeptides are cyclic peptides formed by intramolecular disulfide linking of cysteine residues at both ends of the peptide. The chiral catalysts were constructed by binding the cyclic peptide with Cu2+ ions. The advantage of this system is smallsequence constriction and flexibility in the amino acids of the polypeptides. In catalyzing Diels−Alder and Friedel−Crafts reactions, these cyclic-peptide-based chiral catalysts achieved high enantioselectivities of up to 99% ee and 86% ee, respectively (Figure 156). Furthermore, in this work, Herrmann et al. also

catalysts have totally different supramolecular chirality (Figure 153). In the case of Friedel−Crafts reactions and Michael additions using many different substrates, enantiomers of the products can be obtained by the catalysis of L-DNA- or D-DNAbased supramolecular chiral catalysts.439 In the construction of DNA-based chiral catalysts, even though metal ions are very important, they are not always necessary. Andréasson et al. studied the asymmetric closing reaction of dithienylethene derivatives by complexion with DNA upon photoirradiation. In this study, fluorinated dithienylethene derivatives containing methylpyridinium and methylquinolium substituents were bound to DNA in both the open and the closed forms. Cyclization of dithienylethene derivatives upon photoirradiation could produce the closed form of these molecules with significant enantioselectivity (Figure 154). In this case, chirality was transferred from the DNA to the products.440

Figure 154. Enantioselective cyclization of photochromic dithienylethenes bound to DNA. Reprinted with permission from ref 440. Copyright 2013 John Wiley & Sons.

Although the most popular natural chiral catalysts (enzymes) are polypeptides, constructing artificial enzymes via the coassembly of polypeptides with other molecules has been only partially successful. Polypeptides are generally much more complicated biomacromolecules than DNA, due to the greater array of molecular building blocks available for the formation of polypeptides and the relatively weak but complicated non-

Figure 156. (a) Cyclic peptide ligand with constrained conformation through an intramolecular disulfide bridge. (b) D−A reaction catalyzed by cyclic peptide ligand and Cu2+. Reprinted with permission from ref 442. Copyright 2014 John Wiley & Sons.

Figure 155. Enantioselective artificial metalloenzymes based on a bovine pancreatic polypeptide scaffold, showing catalytic Diels−Alder and Michael addition reactions in water with high enantioselectivities. Reprinted with permission from ref 441. Copyright 2009 John Wiley & Sons. 7380

DOI: 10.1021/cr500671p Chem. Rev. 2015, 115, 7304−7397

Chemical Reviews

Review

crystals, which formed from the assembly of coronene, are at the core of the matter. Thus, stacking at certain twisting angles within the discotic liquid crystals increased charge carrier mobility of these devices.454

systematically studied the relationship between amino acid sequences and the corresponding enantioselectivities of the catalytic reactions. The results show that the position of alanine within the sequences plays a very important role.442 6.4. Optics and Electronics Based on Supramolecular Chiral Assembly

Development of functional devices is one of the main objectives of research on supramolecular assembly. In the application of supramolecular chirality from self-assembled systems, chiral electronic or optical devices are also worth discussing. Recently, these issues have attracted increasing interest.443−448 Chiral optics and electronics are directly dependent on many different functions of supramolecular chiral assemblies. For example, chiral sensors can be used to construct chiral electronic devices, as shown by Wei and co-workers. They prepared ultraordered superhelical microfibers with clear screws and favorable monodispersity from chiral polyaniline (PANI). When these superhelical microfibers were treated with chiral aminohexane vapor with different handedness, very different electrical conductivity in these microfibers was detected (Figure 156).449,450

Figure 157. Superhelical conducting microfibers with homochirality for enantioselective sensing. Reprinted with permission from ref 449. Copyright 2013 American Chemical Society.

Figure 158. (A) Molecular structure of coronene (168). (B) Charge mobilities as a function of temperature as measured by the pulse radiolysis time-resolved microwave conductivity (PR-TRMC) technique. Reprinted with permission from ref 454. Copyright 2009 Nature Publishing Group.

In the development of chiral electronic or optic devices, the most important aspect is the relationship between the optical/ electrical properties of the materials and the supramolecular chirality. The first important work concerning chiral optics was published by Verbiest et al. in 1998. The authors prepared LB films of helicene with different supramolecular chirality and studied the second-order nonlinear optical (NLO) properties of these LB films. The results show that the second-order NLO susceptibility of the chiral assemblies can be 30 times larger than that of the racemic material with the same chemical structure.451 Except for optical properties, the electrical properties of some supramolecular assemblies were also found to be dependent on the material’s chirality. For example, Fourmigué et al. studied the electronic conductivity of chiral salts of tetrathiafulvalene methyl−oxazoline derivatives. The results showed that the conductivity of the pure enantiomeric salts can be an order of magnitude higher than the conductivity of the racemic salts.452 Wei et al. fabricated hierarchical chiral assemblies of the conducting polyaniline (PANI) with different nanostructures and superstructures by controlling the interactions between molecules. The anisotropic electrical transport properties based on the arrangement of molecules and nanostructures were probed.453 The semiconductor properties of supramolecular assemblies has always been very interesting. Remarkably, supramolecular chirality was also found to play a very important role in this issue. Müllen et al. studied field-effect transistor devices based on the assembly of coronene (168). In this system, discotic liquid

The photocurrent properties of the supramolecular chiral assemblies formed at the air/water interface were investigated by the Liu group. They found that an anthracene derivative (AN) could be controllably assembled to nanocoils and straight nanoribbons on water surfaces depending on the different surface pressures. Most interestingly, the nanoribbons exhibited a switchable photocurrent, while the nanocoils did not show a photocurrent response (Figure 159).455 Not only can a photocurrent be generated from the chiral supramolecular assemblies, devices based on chiral supramolecular assemblies can also be used as sensors for detecting circularly polarized light. These results were reported by Fuchter and Campbell et al. In this study, they constructed organic field effect transistors from the assembly of helicene (Figure 160A), and a highly specific photoresponse to circularly polarized light was detected. Importantly, the photoresponse to circularly polarized light was found to be directly related to the handedness of the helicene molecule (Figure 160B).456 6.5. Circularly Polarized Luminescence (CPL) Based on Chiral Supramolecular Assemblies

Circularly polarized light is inherently chiral and has been regarded as one possible origin of natural homochirality457 and the source of chiral information during the emergence of life.308,309 As we described previously, supramolecular chirality with controlled handedness can be introduced into the 7381

DOI: 10.1021/cr500671p Chem. Rev. 2015, 115, 7304−7397

Chemical Reviews

Review

Figure 159. Controllable fabrication of supramolecular nanocoils and nanoribbons and their morphology-dependent photoswitching. Reprinted with permission from ref 455. Copyright 2009 American Chemical Society.

Akagi et al. synthesized chiral polythiophenes and chiral thiophene−phenylene copolymers and found that these polymers, with different molecular structures and aggregation states, could exhibit red, green, and blue fluorescent. Remarkably, mixing these different fluorescent polymers generated a unique, circularly polarized white luminescence.465

Figure 160. (A) Molecular structure and device architecture of the circularly polarized light-detecting helicene OFETs. (B) Response of helicene OFETs to circularly polarized light. Reprinted with permission from ref 456. Copyright 2013 Nature Publishing Group.

supramolecular assemblies containing only achiral molecular building blocks via irradiation with circularly polarized light (CPL). Among the different circularly polarized light, the circularly polarized luminescence from chiral assemblies, abbreviated CPL, can be extremely important and is attracting increasing research interest. In the generation of circularly polarized luminescence (CPL) from chiral supramolecular assemblies, the chiral arrangement of the luminescent chromophores is an essential prerequisite. When luminophores exist in a dissymmetric environment within the photoexcited state, circularly polarized luminescence (CPL) is generated. The study of the circularly polarized luminescence (CPL) from chiral assemblies has been widely dominated by lanthanide complexes owing to their ability to exhibit high CPL dissymmetry.458,459 On the other hand, many different chiral supramolecular assemblies resulting from organic molecular building blocks have also been recently found to be very important sources for generating circularly polarized luminescence (CPL). This situation can be another important application of chiral supramolecular assemblies. The most prominent efforts are based on the π-conjugated polymers with chiral side chains or helical aggregated nanostructures that have been reported to show intense CPL signals.460−465 For example, Swager et al. synthesized chiral poly(pphenylenevinylene) derivatives and studied the circularly polarized luminescence (CPL) spectroscopy of the assemblies from this polymer. Interestingly, using the same polymer with same molecular chirality, different supramolecular assemblies were found to produce opposite CPL spectra.463

Figure 161. Mixture of red, green, and blue fluorescent polymers generated a unique, circularly polarized white luminescence. Reprinted with permission from ref 465. Copyright 2012 American Chemical Society.

Another type of important chiral supramolecular system for generating circularly polarized luminescence is the assembly of helicenes. The aggregation of chiral helicenes was found to show large CPL dissymmetry owing to the strong helical distortion of π systems.466−469 Maeda et al. introduced the BINOL−boron moiety to dipyrrolyldiketones and fabricated the chiral conformation of π-conjugated system. The anions-triggered strong circularly polarized luminescence (CPL) was observed from these assemblies.468 In addition, Nakashima and Kawai et al. reported chiral bichromophoric perylene bisimides as active materials for circularly polarized emission. They found that the compounds formed chiral aggregates with solvent variations. A large enhancement in the dissymmetry of circularly polarized luminescence was achieved by the aggregated structures. It was further found that the spacer between the chiral center and the 7382

DOI: 10.1021/cr500671p Chem. Rev. 2015, 115, 7304−7397

Chemical Reviews

Review

1,4-benzenedicarboxamide phenylalanine derivative (170) as supramolecular gelators to construct different supramolecular hydrogels, and the cell adhesion within these supramolecular hydrogels was studied. It was found that cell adhesion and proliferation can be influenced by the chirality of the nanofibers. Thus, the left-handed helical nanofibers increased cell adhesion and proliferation, while the right-handed nanofibers decrease cell adhesion and proliferation. The stereospecific interaction between chiral nanofibers and fibronectin plays a critical role in these effects (Figure 163).476

chromophoric units played a crucial role in the effective enhancement of chiroptical properties in these self-assembled structures.470

Figure 163. Schematic representation of the culture of cells in supramolecular hydrogels and the different cell-adhesion and cellproliferation behavior in the enantiomeric nanofibrous hydrogels (d: right-handed helical nanofibers; l: left-handed helical nanofibers). The molecular structure of the gelator enantiomers (170) is shown. Reprinted with permission from ref 476. Copyright 2014 John Wiley & Sons.

Figure 162. Enhancement in the dissymmetry of circularly polarized luminescence from the assembly of chiral bichromophoric perylene bisimides. Reprinted with permission from ref 470. Copyright 2013 John Wiley & Sons.

6.6. Biological Applications of Supramolecular Chirality

The existence of life and biological evolution directly depend on both molecular chirality and supramolecular chirality. This situation can be demonstrated from the homochirality of amino acids, nucleic acids, and many other biomolecules as well as the helical nanostructures formed by the folding of DNA and proteins. Many biomedical applications are also closely related to molecular chirality and supramolecular chirality. For example, most drugs used for clinical application are chiral molecules. In general, every aspect of biological applications is dependent on chirality. We cannot address all of these issues in this review. For a further understanding of supramolecular chirality from selfassembled systems, we will discuss some aspects of surpramolecular chirality effects on cell adhesion. Supramolecular hydrogels formed from the self-assembly of peptide derivatives or nucleic acid derivatives have been studied for different biological applications.471−473 Certainly, the chirality always plays a very important role. For example, for the hydrogels formed by short peptides, L-peptides have been found to be labile to proteases.474 Marchesan et al. recently studied the effects of amino acid chirality on tripeptide selfassembly and hydrogelation at physiological pH and cytocompatibility in fibroblast cell culture. In this study, different uncapped hydrophobic tripeptides with all combinations of Dand L-amino acids were prepared. The self-assembly and hydrogelation was found to be dependent on the chirality of the amino acids, and combinations of D, L-amino acids are very useful for maintaining the viability and proliferation of fibroblasts in vitro.475 Interestingly, the cell adhesion in the supramolecular hydrogels was found to be dependent on the handedness of the selfassembled nanofibers. Feng et al. used the two enantiomers of a

As mentioned above, the handedness of self-assembled nanostructures can influence cell proliferation. However, Zouani et al. demonstrated that helical nanostructures with the same handedness, but different shapes and periodicities show totally different capabilities for inducing human mesenchymal stem cell (hMSCs) adhesion and commitment into osteoblast lineage. In this study, mineralization of helical organic nanoribbons, which formed from the self-assembly of Gemini-type amphiphiles, could produce chiral silica nanoribbons with two different shapes and periodicities. Interestingly, helical silica nanoribbons with a specific periodicity of 63 nm (±5 nm) helped the specific cell adhesion and stem cell differentiation, while silica twists with a specific periodicity of 100 nm (±15 nm) did not (Figure 164). These results indicate that stem cells could interpret helical nanostructures with supramolecular chirality.477 Recently, Liu et al. synthesized gelators bearing amphiphilic Lglutamide and D- or L-pantolactone (abbreviated as DPLG and LPLG, 171). The self-assembly of DPLG and LPLG produced nanostructures with opposite supramolecular chirality. The ability of proteins to adhere to these nanostructures was found to be dependent on their supramolecular chirality, as demonstrated from quartz crystal microbalance measurements. Thus, the supramolecular nanostructures formed by DPLG have stronger adhesive ability to human serum albumin. Interestingly, the distinction of protein adhesion ability was only found at the supramolecular level. At the molecular level, however, no clear difference could be detected (Figure 163).478 7383

DOI: 10.1021/cr500671p Chem. Rev. 2015, 115, 7304−7397

Chemical Reviews

Review

Supramolecular chirality can be produced in systems containing chiral and/or achiral molecules. In contrast to the molecular chirality, it is still difficult to quantitatively evaluate the purity of supramolecular chirality. In a system containing chiral molecules, the main questions are do they form only one kind of supramolecular chirality or does there exist a percentage of assemblies with the opposite chirality? Occasionally, supramolecular chirality has emerged from assemblies based on achiral molecules. Even though one can observe microscopic chirality, two enantiomers coexisted. Understanding the emergence of chirality and how to evaluate the enantiomeric excess of a chiral assembly remains difficult. Supramolecular chirality is generally dynamic and strongly related to the self-assembly process. While we achieved many controls over supramolecular chirality, the characterization techniques of the dynamic processes of the chiral assemblies in particular, the development of timedependent spectroscopy and imaging technology is urgently necessary. Although we acquired much knowledge about supramolecular chirality in self-assembly systems related to intermolecular interactions, structural control, and function development, many of these are limited to one to two components. There is a lack of understanding of how to tune many different molecules into a complex chiral system in a cooperative or syndetic way as is accomplished in a living cell. Nanostructured chiral materials offer many opportunities to develop entirely new functional materials, which justifies research into supramolecular chirality. In this regard, can new catalytic, optical, opto-electrical, and magnetic materials result from work on chiral self-assembly systems? Chirality effects are fundamental to biological systems, such as different enantiomers that can be a useful drug and or poisonous depending on their chirality. How to construct the chiral/biointerface? Therefore, further efforts on supramolecular chirality research should integrate new ideas from supramolecular chemistry, biology, medical science, pharmacology, and material and nanosciences.

Figure 164. SEM images of helical silica nanoribbons and silica twists; adhesion and differentiation of stem cells on helical silica nanoribbon substrates. Reprinted with permission from ref 477. Copyright 2013 American Chemical Society.

Figure 165. Self-assembled nanostructures with opposite supramolecular chirality showing different adhesive ability to human serum albumin. Reprinted with permission from ref 478. Copyright 2014 American Chemical Society.

AUTHOR INFORMATION

7. CONCLUSIONS Chiral self-assembly from the molecular to the supramolecular level represents one of the most attractive and promising areas in supramolecular chemistry and self-assembly. The supramolecular chirality in these self-assembled systems is the expression of the noncovalent interactions between the component molecules, where chiral transfer from a chiral component to the whole assembly plays an important role. In addition, supramolecular chirality can also emerge through symmetry breaking even when only achiral molecules are involved. Due to the dynamic features of the self-assembly system, the supramolecular chirality can be regulated through the design of the chiral molecules themselves, external conditions such as pH, metal ions, photoirradiation, solvents, temperature, sonication, and so on. Different from molecular chirality, supramolecular chirality can exhibit unique properties such as the sergeant-and-soldier principle, the majority-rule principle, and chiral memories in several systems. Supramolecular chirality in self-assembled systems has been found to be useful in chiral sensing, chiral molecular recognition, and asymmetric catalysis. Some new functions such as chiroptical switching, chiroptics, and CPL have also been observed. Furthermore, chiral nanostructures showed some interesting properties when interacting with the biological systems. This review has described many examples of the emergence, regulation, and unique features or functions of the supramolecular chirality; however, there are still many unknowns related to supramolecular chirality.

Corresponding Author

*Phone: +86 10 82615803. E-mail: [email protected]. Notes

The authors declare no competing financial interest. Biographies

Minghua Liu, born in 1965 in China, is a Professor at the Institute of Chemistry of the Chinese Academy of Sciences (CAS). He graduated from Nanjing University in 1986 and received his Ph.D. degree in 1994 in Materials Science from Saitama University, Japan, under the supervision of Prof. Kiyoshige Fukuda. He then joined the Institute of 7384

DOI: 10.1021/cr500671p Chem. Rev. 2015, 115, 7304−7397

Chemical Reviews

Review

REFERENCES

Physical and Chemical Research (RIKEN) as a Special Postdoctoral Researcher from 1994 to 1997. He joined the Institute of Photographic Chemistry, CAS, in 1998 and then the Institute of Chemistry, CAS, from 1999. His research interests cover the colloid and interface sciences, selfassembly, supramolecular chemistry, and soft materials, particularly the chirality problems in those systems including monolayers, Langmuir− Blodgett films, supramolecular gels, and soft nanomaterials.

(1) Hegstrom, R. A.; Kondepudi, D. K. The Handedness of the Universe. Sci. Am. 1990, 262, 108−115. (2) Bai, C.; Liu, M. From Chemistry to Nanoscience: Not Just a Matter of Size. Angew. Chem., Int. Ed. 2013, 52, 2678−2683. (3) Lehn, J. M. Supramolecular Chemistry. Concepts and Perspectives; VCH: New York, 1995. (4) In Comprehensive Supramolecular Chemistry; Atwood, J. L., Davies, J. E. D., MacNicol, D. D., Vögtle, F., Lehn, J. M., Eds.; Pergamon Press: Oxford, 1996. (5) Whitesides, G. M.; Boncheva, M. Beyond Molecules: SelfAssembly of Mesoscopic and Macroscopic Components. Proc. Natl. Acad. Sci., U.S.A. 2002, 99, 4769−4774. (6) Alberts, B.; Bray, D.; Lewis, J.; Raff, M.; Roberts, K.; Watson, J. D. Molecular Biology of the Cell; Garland Press: New York, 1994. (7) Feringa, B. L.; van Delden, R. A.; Koumura, N.; Geertsema, E. M. Chiroptical Molecular Switches. Chem. Rev. 2000, 100, 1789−1816. (8) Pérez-García, L.; Amabilino, D. B. Spontaneous Resolution under Supramolecular Control. Chem. Soc. Rev. 2002, 31, 342−356. (9) Mateos-Timoneda, M. A.; Crego-Calama, M.; Reinhoudt, D. N. Supramolecular Chirality of Self-Assembled Systems in Solution. Chem. Soc. Rev. 2004, 33, 363−372. (10) Pérez-García, L.; Amabilino, D. B. Spontaneous Resolution, Whence and Whither: From Enantiomorphic Solids to Chiral Liquid Crystals, Monolayers and Macro-and Supra-Molecular Polymers and Assemblies. Chem. Soc. Rev. 2007, 36, 941−967. (11) Hembury, G. A.; Borovkov, V. V.; Inoue, Y. Chirality-Sensing Supramolecular Systems. Chem. Rev. 2008, 108, 1−73. (12) Amabilino, D. B.; Veciana, J. Supramolecular Chiral Functional Materials; Springer: Berlin, Heidelberg, 2006; p 253. (13) Wang, Y.; Xu, J.; Wang, Y.; Chen, H. Emerging Chirality in Nanoscience. Chem. Soc. Rev. 2013, 42, 2930−2962. (14) De Feyter, S.; De Schryver, F. C. Two-Dimensional Supramolecular Self-Assembly Probed by Scanning Tunneling Microscopy. Chem. Soc. Rev. 2003, 32, 139−150. (15) Ernst, K. H. Supramolecular surface chirality; Springer: Berlin, Heidelberg, 2006, 209. (16) Elemans, J. A. A. W.; De Cat, I.; Xu, H.; De Feyter, S. TwoDimensional Chirality at Liquid-Solid Interfaces. Chem. Soc. Rev. 2009, 38, 722−736. (17) Du, D. Y.; Yan, L. K.; Su, Z. M.; Li, S. L.; Lan, Y. Q.; Wang, E. B. Chiral Polyoxometalate-Based Materials: From Design Syntheses to Functional Applications. Coord. Chem. Rev. 2013, 257, 702−717. (18) Serrano, J. L.; Sierra, T. Helical Supramolecular Organizations from Metal-Organic Liquid Crystals. Coord. Chem. Rev. 2003, 242, 73− 85. (19) Maeda, C.; Kamada, T.; Aratani, N. Chiral Self-discriminative Selfassembling of Meso Linked Diporphyrins. Coord. Chem. Rev. 2007, 251, 2743−2752. (20) Crassous, J. Chiral Transfer in Coordination Complexes: Towards Molecular Materials. Chem. Soc. Rev. 2009, 38, 830−840. (21) Constable, E. C. Stereogenic Metal Centres-From Werner to Supramolecular Chemistry. Chem. Soc. Rev. 2013, 42, 1637−1651. (22) Seeber, G. T. Tiedemann, B. E. F.; Raymond, K. N. Supramolecular Chirality in Coordination Chemistry, Supramolecular Chirality; Springer: Berlin, Heidelberg, 2006, 147. (23) Valev, V. K.; Baumberg, J. J.; Sibilia, C.; Verbiest, T. Chirality and Chiroptical Effects in Plasmonic Nanostructures: Fundamentals, Recent Progress, and Outlook. Adv. Mater. 2013, 25, 2517−2534. (24) Nonaka, K.; Yamaguchi, M.; Yasui, M. Guest-Induced Supramolecular Chirality in a Ditopic Azoprobe-Cyclodextrin Complex in Water. Chem. Commun. 2014, 26, 10059−10061. (25) Joseph, G. Stereochemical Vocabulary for Structures that are Chiral but not Asymmetric: History, Analysis, and Proposal for a Rational Terminology. Chirality 2011, 23, 647−659. (26) Moss, G. P. Basic Terminology of Stereochemistry. Pure Appl. Chem. 1996, 68, 2193−2222. (27) Jiang, J.; Wang, T.; Liu, M. Creating Chirality in The Inner Walls of Silica Nanotubes Through a Hydrogel Template: Chiral Tran-

Li Zhang received her B.S. degree in Physical Chemistry from Shandong University in China (1998) and Ph.D. degree in Physical Chemistry from the Institute of Chemistry of the Chinese Academy of Sciences (2004) under the supervision of Prof. Minghua Liu. Following graduation, she worked for 2 years at Tohoku University as a Postdoctoral Fellow. She has carried out research into chiral supramolecular assemblies formed by achiral porphyrins and asymmetric catalysis of chiral assemblies. She is currently an Associate Professor at the Institute of Chemistry of the Chinese Academy of Sciences.

Tianyu Wang received his B.Sc. and M.Sc. degrees in Chemistry from Tianjin University, China. He received his Ph.D. degree in Organic Chemistry from the Institute of Chemistry of the Chinese Academy of Sciences in 2001. After that he worked as a Postdoctoral Researcher at the Free University of Berlin in Germany with Prof. Dr. J.-H. Fuhrhop. Since 2007 he has been working as an Associate Professor at the Institute of Chemistry, CAS. His research interests are supramolecular assemblies and soft matters.

ACKNOWLEDGMENTS This work was supported by the Basic Research Development Program (2013CB834504), the National Natural Science Foundation of China (Nos. and 21321063, 21473219, 21474118, and 91427302), and the Fund of the Chinese Academy of Sciences (No. XDB12020200). 7385

DOI: 10.1021/cr500671p Chem. Rev. 2015, 115, 7304−7397

Chemical Reviews

Review

scription and Chiroptical Switch. Chem. Commun. 2010, 46, 7178− 7180. (28) Weckesser, J.; De Vita, A.; Barth, J. V.; Cai, C.; Kern, K. Mesoscopic Correlation of Supramolecular Chirality in One-Dimensional Hydrogen-Bonded Assemblies. Phys. Rev. Lett. 2001, 87, 096101. (29) Rong, Y.; Chen, P.; Liu, M. Self-Assembly of Water-Soluble TPPS in Organic Solvents: From Nanofibers to Mirror Imaged Chiral Nanorods. Chem. Commun. 2013, 49, 10498−10500. (30) Duan, P.; Zhu, X.; Liu, M. Isomeric Effect in the Self-Assembly of Pyridine-Containing L-Glutamic Lipid: Substituent Position Controlled Morphology and Supramolecular Chirality. Chem. Commun. 2011, 47, 5569−5571. (31) Chen, J. C. Introduction to Scanning Tunneling Microscopy; Oxford University Press: New York, 1993. (32) Xu, Q.; Wang, D.; Wan, L.; Wang, C.; Bai, C.; Feng, G.; Wang, M. Discriminating Chiral Molecules of (R)-PPA and (S)-PPA in Aqueous Solution by ECSTM. Angew. Chem., Int. Ed. 2002, 41, 3408−3411. (33) In Circular Dichroism Principles and Applications; Berova, N., Nakanishi, K., Woody, R., Eds.; John Wiley & Sons: New York, 2000. (34) Taniguchi, T.; Usuki, T. Circular Dichroism Spectroscopy. Supramolecular Chemistry:From Molecules to Nanomaterials; John Wiley & Sons, Ltd.: New York, 2012. (35) Polavarapu, P. L. Ab Initio Vibrational Raman and Raman Optical Activity Spectra. J. Phys. Chem. 1990, 94, 8106−8112. (36) Fischer, P.; Hache, F. Nonlinear Optical Spectroscopy of Chiral Molecules. Chirality 2005, 17, 421−437. (37) Gottarelli, G.; Lena, S.; Masiero, S.; Pieraccini, S.; Spada, G. P. The Use of Circular Dichroism Spectroscopy for Studying the Chiral Molecular Self-assembly: An Overview. Chirality 2008, 20, 471−485. (38) Spitz, C.; Dähne, S.; Ouart, A.; Abraham, H. W. Proof of Chirality of J-Aggregates Spontaneously and Enantioselectively Generated from Achiral Dyes. J. Phys. Chem. B 2000, 104, 8664−8669. (39) Yuan, J.; Liu, M. Chiral Molecular Assemblies from a Novel Achiral Amphiphilic 2-(Heptadecyl) Naphtha[2,3]imidazole through Interfacial Coordination. J. Am. Chem. Soc. 2003, 125, 5051−5056. (40) Arteaga, O.; Canillas, A.; Purrello, R.; Ribó, J. M. Evidence of Induced Chirality in Stirred Solutions of Supramolecular Nanofibers. Opt. Lett. 2009, 34, 2177−2179. (41) Danila, I.; Pop, F.; Escudero, C.; Feldborg, L. N.; Puigmart-Luis, J.; Riobe, F.; Avarvari, N.; Amabilino, D. B. Twists and Turns in the Hierarchical Self-Assembly Pathways of a Non-Amphiphilic Chiral Supramolecular Material. Chem. Commun. 2012, 4552−4554. (42) Kunitake, T.; Kimizuka, N.; Higashi, N.; Nakashima, N. BilayerMembranes of Triple-Chain Ammonium Amphiphiles. J. Am. Chem. Soc. 1984, 106, 1978−1983. (43) Nakashima, N.; Asakuma, S.; Kunitake, T. Optical Microscopic Study of Helical Superstructuresof Chiral Bilayer Membranes. J. Am. Chem. Soc. 1985, 107, 509−510. (44) Sorrenti, A.; Illa, O.; Ortuño, R. M. Amphiphiles in Aqueous Solution: Well Beyond a Soap Bubble. Chem. Soc. Rev. 2013, 42, 8200− 8219. (45) Yui, H.; Minamikawa, H.; Danev, R.; Nagayama, K.; Kamiya, S.; Shimizu, T. Growth Process and Molecular Packing of a Self-assembled Lipid Nanotube: Phase-Contrast Transmission Electron Microscopy and XRD Analyses. Langmuir 2008, 24, 709−713. (46) Kamiya, S.; Minamikawa, H.; Jung, J. H.; Yang, B.; Masuda, M.; Shimizu, T. Molecular Structure of Glucopyranosylamide Lipid and Nanotube Morphology. Langmuir 2005, 21, 743−750. (47) Barclay, T. G.; Constantopoulos, K.; Zhang, W.; Fujiki, M.; Petrovsky, N.; Matisons, J. G. Chiral Self-Assembly of Designed Amphiphiles: Influences on Aggregate Morphology. Langmuir 2013, 29, 10001−10010. (48) Fuhrhop, J. H.; Helfrich, W. Fluid and Solid Fibers Made of Lipid Molecular Bilayers. Chem. Rev. 1993, 93, 1565−1582. (49) Brizard, A.; Oda, R.; Huc, I. In Low Molecular Mass Gelators. Design, Self-Assembly, Function; Fages, F., Ed., Springer-Verlag: Berlin, Heidelberg, 2005; Vol. 256.

(50) Shimizu, T.; Masuda, M.; Minamikawa, H. Supramolecular Nanotube Architectures Based on Amphiphilic Molecules. Chem. Rev. 2005, 105, 1401−1443. (51) Li, Y.; Wang, T.; Liu, M. Gelating-Induced Supramolecular Chirality of Achiral Porphyrins: Chiroptical Switch between Achiral Molecules and Chiral Assemblies. Soft Matter 2007, 3, 1312−1317. (52) Li, Y.; Liu, M. Fabrication of Chiral Silver Nanoparticles and Chiral Nanoparticulate Film via Organogel. Chem. Commun. 2008, 5571−5573. (53) Zhu, X.; Li, Y.; Duan, P.; Liu, M. Self-Assembled Ultralong Chiral Nanotubes and Tuning of Their ChiralityThrough the Mixing of Enantiomeric Components. Chem.Eur. J. 2010, 16, 8034−8040. (54) Miao, W.; Zhang, L.; Wang, X.; Cao, H.; Jin, Q.; Liu, M. A DualFunctional Metallogel of Amphiphilic Copper(II) Quinolinol:Redox Responsiveness and Enantioselectivity. Chem.Eur. J. 2013, 19, 3029− 3036. (55) Miao, W.; Zhang, L.; Wang, X.; Qin, L.; Liu, M. Gelation-Induced Visible Supramolecular Chiral Recognition by Fluorescent Metal Complexes of Quinolinol−Glutamide. Langmuir 2013, 29, 5435−5442. (56) Duan, P.; Li, Y.; Li, L.; Deng, J.; Liu, M. Multiresponsive Chiroptical Switch of an Azobenzene-ContainingLipid: Solvent, Temperature, and Photoregulated Supramolecular Chirality. J. Phys. Chem. B 2011, 115, 3322−3329. (57) Wang, X.; Duan, P.; Liu, M. Universal Chiral Twist via Metal Ion Induction in the Organogel of Terephthalic Acid Substituted Amphiphilic L-Glutamide. Chem. Commun. 2012, 48, 7501−7503. (58) Jin, Q.; Zhang, L.; Liu, M. Solvent-Polarity-Tuned Morphology and Inversion of SupramolecularChirality in a Self-Assembled Pyridylpyrazole-Linked Glutamide Derivative:Nanofibers, Nanotwists, Nanotubes, and Microtubes. Chem.Eur. J. 2013, 19, 9234−9241. (59) Wang, X.; Duan, P.; Liu, M. Organogelation-Controlled Topochemical [2 + 2] Cycloaddition andMorphological Changes: From Nanofiber to Peculiar Coaxial HollowToruloid-Like Nanostructures. Chem.Eur. J. 2013, 19, 16072−16079. (60) Zhang, L.; Liu, C.; Jin, Q.; Zhu, X.; Liu, M. Pyrene-Functionalized Organogel and Spacer Effect: From Emissive Nanofiber to Nanotube and Inversion of Supramolecular Chirality. Soft Matter 2013, 9, 7966− 7973. (61) Liu, C.; Jin, Q.; Lv, K.; Zhang, L.; Liu, M. Water Tuned the Helical Nanostructures andSupramolecular Chirality in Organogels. Chem. Commun. 2014, 50, 3702−3705. (62) Wang, X.; Liu, M. Vicinal Solvent Effect on Supramolecular Gelation: Alcohol Controlled Topochemical Reaction and the Toruloid Nanostructure. Chem.Eur. J. 2014, 20, 10110−10116. (63) Takafuji, M.; Kira, Y.; Tsuji, H.; Sawada, S.; Hachisako, H.; Ihara, H. Optically Active Polymer Film Tuned by a Chirally Self-Assembled Molecular Organogel. Tetrahedron 2007, 63, 7489−7494. (64) Shirosaki, T.; Chowdhury, S.; Takafuji, M.; Alekperov, D.; Popova, G.; Hachisako, H.; Ihara, H. Functional Organogels from Lipophilic L-Glutamide Derivative Immobilized on Cyclotriphosphazene Core. J. Mater. Res. 2006, 21, 1274−1278. (65) Kira, Y.; Okazaki, Y.; Sawada, T.; Takafuji, M.; Ihara, H. Amphiphilic Molecular Gels from Omega-Aminoalkylated L-Glutamic Acid Derivatives with Unique Chiroptical Properties. Amino Acids 2010, 39, 587−597. (66) Jintoku, H.; Sagawa, T.; Sawada, T.; Takafuji, M.; Hachisako, H.; Ihara, H. Molecular Organogel-Forming Porphyrin Derivative with Hydrophobic L-Glutamide. Tetrahedron Lett. 2008, 49, 3987−3990. (67) Oda, R.; Huc, I.; Schmutz, M.; Candau, S. J.; MacKintosh, F. C. Tuning Bilayer Twist Using Chiral Counterions. Nature 1999, 399, 566−569. (68) Berthier, D.; Buffeteau, T.; Léger, J.-M.; Oda, R.; Huc, I. From Chiral Counterions to Twisted Membranes. J. Am. Chem. Soc. 2002, 124, 13486−13494. (69) Brizard, A.; Aimé, C.; Labrot, T.; Huc, I.; Berthier, D.; Artzner, F.; Desbat, B.; Oda, R. Counterion, Temperature, and Time Modulation of Nanometric Chiral Ribbons from Gemini-Tartrate Amphiphiles. J. Am. Chem. Soc. 2007, 129, 3754−3762. 7386

DOI: 10.1021/cr500671p Chem. Rev. 2015, 115, 7304−7397

Chemical Reviews

Review

(70) Fuhrhop, A. H.; Wang, T. Bolaamphiphiles. Chem. Rev. 2004, 104, 2901−2937. (71) Shimizu, T.; Kogiso, M.; Masuda, M. Vesicle Assembly in Microtubes. Nature 1996, 383, 487−488. (72) Zhan, C.; Gao, P.; Liu, M. Self-Assembled Helical SphericalNanotubes from an L-Glutamic Acid Based Bolaamphiphilic Low Molecular Mass Organogelator. Chem. Commun. 2005, 462−464. (73) Wang, T.; Jiang, J.; Liu, Y.; Li, Z.; Liu, M. Hierarchical SelfAssembly of Bolaamphiphiles with a Hybrid Spacer and L-Glutamic Acid Headgroup: pH- and Surface-Triggered Hydrogels, Vesicles, Nanofibers, and Nanotubes. Langmuir 2010, 26, 18694−18700. (74) Liu, Y.; Wang, T.; Li, Z.; Liu, M. Copper(II) Ion Selective and Strong Acid-Tolerable Hydrogels Formed by an L-Histidine Ester Terminated Bolaamphiphile: from Single Molecular Thick Nanofibers to Single-Wall Nanotubes. Chem. Commun. 2013, 49, 4767−4769. (75) Liu, Y.; Wang, T.; Huan, Y.; Li, Z.; He, G.; Liu, M. Self-Assembled Supramolecular Nanotube Yarn. Adv. Mater. 2013, 25, 5875−5879. (76) Masuda, M.; Shimizu, T. Lipid Nanotubes and Microtubes: Experimental Evidence for Unsymmetrical Monolayer Membrane Formation from Unsymmetrical Bolaamphiphiles. Langmuir 2004, 20, 5969−5977. (77) Stupp, S. I.; Zha, R. H.; Palmer, L. C.; Cui, H.; Bitton, R. SelfAssembly of Biomolecular Soft Matter. Faraday Discuss. 2013, 166, 9− 30. (78) Muraoka, T.; Cui, H.; Stupp, S. I. Quadruple Helix Formation of a Photoresponsive Peptide Amphiphile and Its Light-Triggered Dissociation into Single Fibers. J. Am. Chem. Soc. 2008, 130, 2946−2947. (79) Moyer, T. J.; Cui, H.; Stupp, S. I. Tuning Nanostructure Dimensions with Supramolecular Twisting. J. Phys. Chem. B 2013, 117, 4604−4610. (80) Cui, H.; Cheetham, A. G.; Pashuck, E. T.; Stupp, S. I. Amino Acid Sequence in Constitutionally Isomeric Tetrapeptide Amphiphiles Dictates Architecture of One-Dimensional Nanostructures. J. Am. Chem. Soc. 2014, 136, 12461−12468. (81) Duan, P.; Liu, M. Self-Assembly of L-Glutamate Based Aromatic Dendrons through the Air/Water Interface: Morphology, Photodimerization and Supramolecular Chirality. Phys. Chem. Chem. Phys. 2010, 12, 4383−4389. (82) Soininen, A. J.; Kasëmi, E.; Schlüeter, A. D.; Ikkala, O.; Ruokolainen, J.; Mezzenga, R. Self-Assembly and Induced Circular Dichroism in Dendritic Supramolecules with Cholesteric Pendant Groups. J. Am. Chem. Soc. 2010, 132, 10882−10890. (83) Ikeda, M.; Momotake, A.; Kanna, Y.; Shimotori, K.; Arai, T. SelfAssembly of a Poly(Glutamate) Dendrimer: Solvent-Dependent Expression of Molecular Chirality and a 2 + 2 Photocrosslinking Reaction. Photochem. Photobiol. Sci. 2012, 11, 1524−1527. (84) Percec, V.; Imam, M. R.; Peterca, M.; Wilson, D. A.; Graf, R.; Spiess, H. W.; Balagurusamy, V. S. K.; Heiney, P. A. Self-Assembly of Dendronized Triphenylenes into Helical Pyramidal Columns and Chiral Spheres. J. Am. Chem. Soc. 2009, 131, 7662−7677. (85) Rosen, B. M.; Peterca, M.; Morimitsu, K.; Dulcey, A. E.; Leowanawat, P.; Resmerita, A.-M.; Imam, M. R.; Percec, V. Programming the Supramolecular Helical Polymerization of Dendritic Dipeptides via the Stereochemical Information of the Dipeptide. J. Am. Chem. Soc. 2011, 133, 5135−5151. (86) Peterca, M.; Imam, M. R.; Ahn, C. H.; Balagurusamy, V. S.; Wilson, D. A.; Rosen, B. M.; Percec, V. Transfer, Amplification, and Inversion of Helical Chirality Mediated by Concerted Interactions of C3Supramolecular Dendrimers. J. Am. Chem. Soc. 2011, 133, 2311−2328. (87) Duan, P.; Qin, L.; Zhu, X.; Liu, M. Hierarchical Self-Assembly of Amphiphilic Peptide Dendrons: Evolution of Diverse Chiral Nanostructures through Hydrogel Formation over a Wide pH Range. Chem.Eur. J. 2011, 17, 6389−6395. (88) Moberg, C. C3 Symmetry in Asymmetric Catalysis and Chiral Recognition. Angew. Chem., Int. Ed. 1998, 37, 248−268. (89) Gibson, S. E.; Castaldi, M. P. Applications of Chiral C3-Symmetric Molecules. Chem. Commun. 2006, 3045−3062. (90) Brunsveld, L.; Schenning, A.; Broeren, M. A. C.; Janssen, H. M.; Vekemans, J.; Meijer, E. W. Chiral Amplification in Columns of Self-

Assembled N,N′,N″-Tris((S)-3,7-Dimethyloctyl)Benzene-1,3,5-Tricarboxamide In Dilute Solution. Chem. Lett. 2000, 292−293. (91) Smulders, M. M. J.; Buffeteau, T.; Cavagnat, D.; Wolffs, M.; Schenning, A. P. H. J.; Meijer, E. W. C3-Symmetrical Self-Assembled Structures Investigated byVibrational Circular Dichroism. Chirality 2008, 20, 1016−1022. (92) Cantekin, S.; Balkenende, D. W. R.; Smulders, M. M. J.; Palmans, A. R. A.; Meijer, E. W. The Effect of Isotopic Substitution on the Chiralityof a Self-Assembled Helix. Nat. Chem. 2011, 3, 42−46. (93) Nakano, Y.; Hirose, T.; Stals, P. J. M.; Meijer, E. W.; Palmans, A. R. A. Conformational Analysis of Supramolecular Polymerization Processes of Disc-Like Molecules. Chem. Sci. 2012, 3, 148−155. (94) Palmans, A. R. A.; Vekemans, J.; Hikmet, R. A.; Fischer, H.; Meijer, E. W. Lyotropic Liquid-Crystalline Behavior in Disc-Shaped Compounds Incorporating the 3,3′-Di(Acylamino)-2,2′-Bipyridine Unit. Adv. Mater. 1998, 10, 873−876. (95) van Gorp, J. J.; Vekemans, J.; Meijer, E. W. C3-Symmetrical Supramolecular Architectures: Fibers and Organic Gels from Discotic Trisamides and Trisureas. J. Am. Chem. Soc. 2002, 124, 14759−14769. (96) Danila, I.; Riobé, F.; Piron, F.; Puigmarti-Luis, J.; Wallis, J. D.; Linares, M.; Ågren, H.; Beljonne, D.; Amabilino, D. B.; Avarvari, N. C3Symmetrical Supramolecular Architectures: Fibers and Organic Gels from Discotic Trisamides and Trisureas. J. Am. Chem. Soc. 2011, 133, 8344−8353. (97) Narayan, B.; Kulkarni, C.; George, S. J. Synthesis and SelfAssembly of a C3-Symmetric Benzene-1,3,5-Tricarboxamide (BTA) Anchored Naphthalene Diimide Disc. J. Mater. Chem.C 2013, 1, 626− 629. (98) van Hameren, R.; van Buul, A. M.; Castriciano, M. A.; Villari, V.; Micali, N.; Schön, P.; Speller, S.; Scolaro, L. M.; Rowan, A. E.; Elemans, J. A. A. W.; Nolte, R. J. M. Supramolecular Porphyrin Polymers in Solution and at the Solid-Liquid Interface. Nano Lett. 2008, 8, 253−259. (99) Veling, N.; van Hameren, R.; van Buul, A. M.; Rowan, A. E.; Nolte, R. J.; Elemans, J. A. Solvent-Dependent Amplification of Chirality in Assemblies of Porphyrin Trimers Based on Benzene Tricarboxamide. Chem. Commun. 2012, 48, 4371−4373. (100) Cantekin, S.; de Greef, T. F. A.; Palmans, A. R. A. Benzene-1,3,5tricarboxamide: A Versatile Ordering Moiety for Supramolecular Chemistry. Chem. Soc. Rev. 2012, 41, 6125−6137. (101) Stals, P. J.; Everts, J. C.; de Bruijn, R.; Filot, I. A.; Smulders, M. M.; Martin-Rapun, R.; Pidko, E. A.; de Greef, T. F.; Palmans, A. R.; Meijer, E. W. Dynamic Supramolecular Polymers Based on Benzene1,3,5-Tricarboxamides: The Influence of Amide Connectivity on Aggregate Stability and Amplification of Chirality. Chem.Eur. J. 2010, 16, 810−821. (102) Aparicio, F.; García, F.; Sánchez, L. Supramolecular Polymerization of C3-Symmetric Organogelators: Cooperativity, Solvent, and Gelation Relationship. Chem.Eur. J. 2013, 19, 3239−3248. (103) García, F.; Buendia, J.; Sánchez, L. Supramolecular Ribbons from Amphiphilic Trisamides Self-Assembly. J. Org. Chem. 2011, 76, 6271− 6276. (104) García, F.; Sánchez, L. Structural Rules for the Chiral Supramolecular Organization of OPE-based Discotics: Induction of Helicity and Amplification of Chirality. J. Am. Chem. Soc. 2012, 134, 734−742. (105) Wang, F.; Gillissen, M. A. J.; Stals, P. J. M.; Palmans, A. R. A.; Meijer, E. W. Hydrogen Bonding Directed Supramolecular Polymerisation of Oligo(Phenylene-Ethynylene)s: Cooperative Mechanism, Core Symmetry Effect and Chiral Amplification. Chem.Eur. J. 2012, 18, 11761−11770. (106) Cao, H.; Duan, P.; Zhu, X.; Jiang, J.; Liu, M. Self-Assembled Organic Nanotubes through Instant Gelation and Universal Capacity for Guest Molecule Encapsulation. Chem.Eur. J. 2012, 18, 5546−5550. (107) Wang, C.; Dong, H.; Hu, W.; Liu, Y.; Zhu, D. Semiconducting πConjugated Systems in Field-Effect Transistors: A Material Odyssey of Organic Electronics. Chem. Rev. 2012, 112, 2208−2267. (108) Kim, F. S.; Ren, G.; Jenekhe, S. A. One-Dimensional Nanostructures of π-Conjugated Molecular Systems: Assembly, Proper7387

DOI: 10.1021/cr500671p Chem. Rev. 2015, 115, 7304−7397

Chemical Reviews

Review

ties, and Applications from Photovoltaics, Sensors, and Nanophotonics to Nanoelectronics. Chem. Mater. 2011, 23, 682−732. (109) Beaujuge, P. M.; Fréchet, J. M. Molecular Design and Ordering Effects in π-Functional Materials for Transistor and Solar Cell Applications. J. Am. Chem. Soc. 2011, 133, 20009−20029. (110) Gorl, D.; Zhang, X.; Würthner, F. Molecular Assemblies of Perylene Bisimide Dyes in Water. Angew. Chem., Int. Ed. 2012, 51, 6328−6348. (111) Babu, S. S.; Praveen, V. K.; Ajayaghosh, A. Functional piGelators and Their Applications. Chem. Rev. 2014, 114, 1973−2129. (112) Praveen, V. K.; Ranjith, C.; Bandini, E.; Ajayaghosh, A.; Armaroli, N. Oligo(Phenylenevinylene) Hybrids and Self-Assemblies: Versatile Materials for Excitation Energy Transfer. Chem. Soc. Rev. 2014, 43, 4222−4242. (113) Das, R. K.; Kandanelli, R.; Linnanto, J.; Bose, K.; Maitra, U. Supramolecular Chirality in Organogels: A Detailed Spectroscopic, Morphological, and Rheological Investigation of Gels (and Xerogels) Derived from Alkyl Pyrenyl Urethanes. Langmuir 2010, 26, 16141− 16149. (114) Jin, W.; Fukushima, T.; Niki, M.; Kosaka, A.; Ishii, N.; Aida, T. Self-Assembled Graphitic Nanotubes with One-Handed Helical Arrays of A Chiral Amphiphilic Molecular Graphene. Proc. Natl. Acad. Sci., U.S.A, 2005, 102, 10801−10806. (115) Xie, Z.; Stepanenko, V.; Radacki, K.; Würthner, F. Chiral JAggregates of Atropo-Enantiomeric Perylene Bisimides and Their SelfSorting Behavior. Chem.Eur. J. 2012, 18, 7060−7070. (116) Wang, K.; An, H.; Wu, L.; Zhang, J.; Li, X. Chiral Self-Assembly of Lactose Functionalized Perylene Bisimides as Multivalent Glycoclusters. Chem. Commun. 2012, 48, 5644−5653. (117) Bai, S.; Debnath, S.; Javid, N.; Frederix, P. W. J. M.; Fleming, S.; Pappas, C.; Ulijn, R. V. Differential Self-Assembly and Tunable Emission of Aromatic Peptide Bola-Amphiphiles Containing Perylene Bisimide in Polar Solvents Including Water. Langmuir 2014, 30, 7576−7584. (118) Lu, X.; Guo, Z.; Sun, C.; Tian, H.; Zhu, W. Helical Assembly Induced by Hydrogen Bonding from Chiral Carboxylic Acids Based on Perylene Bisimides. J. Phys. Chem. B 2011, 115, 10871−10876. (119) Kumar, J.; Nakashima, T.; Tsumatori, H.; Kawai, T. Circularly Polarized Luminescence in Chiral Aggregates: Dependence of Morphology on Luminescence Dissymmetry. J. Phys. Chem. Lett. 2014, 5, 316−321. (120) Marty, R.; Nigon, R.; Leite, D.; Frauenrath, H. Two-Fold OddEven Effect in Self-Assembled Nanowires from Oligopeptide-PolymerSubstituted Perylene Bisimides. J. Am. Chem. Soc. 2014, 136, 3919− 3927. (121) Chen, S.; Ma, C.; Huang, Z.; Lee, M. Controlled Helicity of the Rigid-Flexible Molecular Assembly Triggered by Water Addition: From Nanocrystal to Liquid Crystal Gel and Aqueous Nanofibers. J. Phys. Chem. C 2014, 118, 8181−8186. (122) Huang, Z.; Kang, S. K.; Banno, M.; Yamaguchi, T.; Lee, D.; Seok, C.; Yashima, E.; Lee, M. Pulsating Tubules from Noncovalent Macrocycles. Science 2012, 337, 1521−1526. (123) Aparicio, F.; Nieto-Ortega, B.; Nájera, F.; Ramírez, F. J.; Navarrete, J. T. L.; Casado, J.; Sánchez, L. Inversion of Supramolecular Helicity in Oligo-p-phenylene-Based Supramolecular Polymers: Influence of Molecular Atropisomerism. Angew. Chem., Int. Ed. 2014, 53, 1373−1377. (124) Helmich, F.; Lee, C. C.; Nieuwenhuizen, M. M. L.; Gielen, J. C.; Christianen, P. C. M.; Larsen, A.; Fytas, G.; Leclere, P. E. L. G.; Schenning, A. P. H. J.; Meijer, E. W. Dilution-Induced Self-Assembly of Porphyrin Aggregates: A Consequence of Coupled Equilibria. Angew. Chem., Int. Ed. 2010, 49, 3939−3942. (125) Maiti, N. C.; Mazumdar, S.; Periasamy, N. J- and H-aggregates of porphyrin-surfactant complexes: Time-resolved fluorescence and other spectroscopic studies. J. Phys. Chem. B 1998, 102, 1528−1538. (126) Borovkov, V. V.; Lintuluoto, J. M.; Inoue, Y. Syn-anti conformational changes in zinc porphyrin dimers induced by temperature-controlled alcohol ligation. J. Phys. Chem. B 1999, 103, 5151− 5156.

(127) Palmans, A. R. A.; Vekemans, J.; Havinga, E. E.; Meijer, E. W. Sergeants-And-Soldiers Principle in Chiral Columnar Stacks of DiscShaped Molecules With C3 Symmetry. Angew. Chem., Int. Ed. 1997, 36, 2648−2651. (128) Wang, Q.; Chen, Y.; Ma, P.; Lu, J.; Zhang, X.; Jiang, J. Morphology and Chirality Controlled Self-Assembled Nanostructures of Porphyrin-Pentapeptide Conjugate: Effect of the Peptide Secondary Conformation. J. Mater. Chem. 2011, 21, 8057−8065. (129) Narayan, B.; Senanayak, S. P.; Jain, A.; Narayan, K. S.; George, S. J. Self-Assembly of π-Conjugated Amphiphiles: Free Standing, Ordered Sheets with Enhanced Mobility. Adv. Funct. Mater. 2013, 23, 3053− 3060. (130) Matmour, R.; De Cat, I.; George, S. J.; Adriaens, W.; Leclere, P.; Bomans, P. H. H.; Sommerdijk, N. A. J. M.; Gielen, J. C.; Christianen, P. C. M.; Heldens, J. T.; van Hest, J. C. M.; Lowik, D. W. P. M.; De Feyter, S.; Meijer, E. W.; Schenning, A. P. H. J. Oligo(p-phenylenevinylene)Peptide Conjugates: Synthesis and Self-Assembly in Solution and at the Solid-Liquid Interface. J. Am. Chem. Soc. 2008, 130, 14576−14583. (131) Kas, O. Y.; Charati, M. B.; Rothberg, L. J.; Galvin, M. E.; Kiick, K. L. Regulation of Electronic Behavior via Confinement of PPV-Based Oligomers on Peptide Scaffolds. J. Mater. Chem. 2008, 18, 3847−3854. (132) Nakatani, Y.; Furusho, Y.; Yashima, E. Synthesis of Helically Twisted 1 + 1 Macrocycles Assisted by Amidinium-Carboxylate Salt Bridges and Control of Their Chiroptical Properties. Org. Biomol. Chem. 2013, 11, 1614−1623. (133) Nisha, S. K.; Asha, S. K. Chiral Poly(L-lactic acid) Driven Helical Self-Assembly of Oligo(p-Phenylenevinylene). J. Mater. Chem. C 2014, 2, 2051−2060. (134) Henze, O.; Feast, W. J.; Gardebien, F.; Jonkheijm, P.; Lazzaroni, R.; Leclère, P.; Meijer, E. W.; Schenning, A. Chiral Amphiphilic SelfAssembled α-α′-Linked Quinque-, Sexi-, and Septithiophenes: Synthesis, Stability and Odd-Even Effects. J. Am. Chem. Soc. 2006, 128, 5923−5929. (135) Digennaro, A.; Wennemers, H.; Joshi, G.; Schmid, S.; MenaOsteritz, E.; Bäuerle, P. Chiral Suprastructures of Asymmetric Oligothiophene-Hybrids Induced by a Single Proline. Chem. Commun. 2013, 49, 10929−10931. (136) Schillinger, E.; Kuemin, M.; Digennaro, A.; Mena-Osteritz, E.; Schmid, S.; Wennemers, H.; Baeuerle, P. Guiding Suprastructure Chirality of an Oligothiophene by a Single Amino Acid. Chem. Mater. 2013, 25, 4511−4521. (137) Nilsson, K. P. R.; Rydberg, J.; Baltzer, L.; Inganäs, O. Twisting Macromolecular Chains: Self-Assembly of a Chiral Supermolecule from Nonchiral Polythiophene Polyanions and Random-Coil Synthetic Peptides. Proc. Natl. Acad. Sci., U.S.A. 2004, 101, 11197−11202. (138) Van den Bergh, K.; Cosemans, I.; Verbiest, T.; Koeckelberghs, G. Expression of Supramolecular Chirality in Block Copoly(thiophene)s. Macromolecules 2010, 43, 3794−3800. (139) Tam, A. Y.; Wong, K. M.; Yam, V. W. Influence of Counteranion on the Chiral Supramolecular Assembly of Alkynylplatinum(II) Terpyridyl Metallogels that Are Stabilised by Pt···Pt and pi-pi Interactions. Chem.Eur. J. 2009, 15, 4775−4778. (140) Leung, S. Y.; Lam, W. H.; Yam, V. W. Dynamic Scaffold of Chiral Binaphthol Derivatives with the Alkynylplatinum(II) Terpyridine Moiety. Proc. Natl. Acad. Sci., U.S.A. 2013, 110, 7986−7991. (141) Pandeeswar, M.; Avinash, M. B.; Govindaraju, T. Chiral Transcription and Retentive Helical Memory: Probing Peptide Auxiliaries Appended with Naphthalenediimides for Their OneDimensional Molecular Organization. Chem.Eur. J. 2012, 18, 4818−4822. (142) Fu, Y.; Li, B.; Huang, Z.; Li, Y.; Yang, Y. Terminal Is Important for the Helicity of the Self-Assemblies of Dipeptides Derived from Alanine. Langmuir 2013, 29, 6013−6017. (143) Imai, Y.; Murata, K.; Nakano, Y.; Harada, T.; Kinuta, T.; Tajima, N.; Sato, T.; Fujiki, M.; Kuroda, R.; Matsubara, Y. Control of Solid-State Chiral Optical Properties of a Chiral Supramolecular Organic Fluorophore Consisting of L-Pyrenesulfonic Acid and Chiral Amine Molecules. Cryst. Eng. Commun. 2010, 12, 1688−1692. 7388

DOI: 10.1021/cr500671p Chem. Rev. 2015, 115, 7304−7397

Chemical Reviews

Review

(144) George, S. J.; Tomovic, Z.; Schenning, A. P.; Meijer, E. W. Insight Into the Chiral Induction in Supramolecular Stacks through Preferential Chiral Solvation. Chem. Commun. 2011, 47, 3451−3453. (145) Nishiguchi, N.; Kinuta, T.; Nakano, Y.; Harada, T.; Tajima, N.; Sato, T.; Fujiki, M.; Kuroda, R.; Matsubara, Y.; Imai, Y. Control of the Solid-State Chiral Optical Properties of a Supramolecular Organic Fluorophore Containing 4-(2-Arylethynyl)-Benzoic Acid. Chem. Asian J. 2011, 6, 1092−1098. (146) George, S. J.; de Bruijn, R.; Tomovic, Z.; Van Averbeke, B.; Beljonne, D.; Lazzaroni, R.; Schenning, A. P. H. J.; Meijer, E. W. Asymmetric Noncovalent Synthesis of Self-Assembled One-Dimensional Stacks by a Chiral Supramolecular Auxiliary Approach. J. Am. Chem. Soc. 2012, 134, 17789−17796. (147) Samanta, S. K.; Bhattacharya, S. Excellent Chirality Transcription in Two-Component Photochromic Organogels Assembled through J-Aggregation. Chem. Commun. 2013, 49, 1425−1427. (148) Shiraki, T.; Dawn, A.; Tsuchiya, Y.; Yamamoto, T.; Shinkai, S. Unexpected Chiral Induction from Achiral Cationic Polythiophene Aggregates and its Application to the Sugar Pattern Recognition. Chem. Commun. 2012, 48, 7091−7099. (149) Zhang, W.; Fujiki, M.; Zhu, X. Chiroptical Nanofibers Generated from Achiral Metallophthalocyanines Induced by Diamine Homochirality. Chem.Eur. J. 2011, 17, 10628−10635. (150) Zhu, X.; Duan, P.; Zhang, L.; Liu, M. Regulation of the Chiral Twist and Supramolecular Chirality in Co-Assemblies of Amphiphilic LGlutamic Acid with Bipyridines. Chem.Eur. J. 2011, 17, 3429−3437. (151) Guo, Z.; De Cat, I.; Van Averbeke, B.; Lin, J.; Wang, G.; Xu, H.; Lazzaroni, R.; Beljonne, D.; Meijer, E. W.; Schenning, A. P.; De Feyter, S. Nucleoside-Assisted Self-Assembly of Oligo(p-phenylenevinylene)s at Liquid/Solid Interface: Chirality and Nanostructures. J. Am. Chem. Soc. 2011, 133, 17764−17771. (152) Guo, Z.; De Cat, I.; Van Averbeke, B.; Ghijsens, E.; Lin, J.; Xu, H.; Wang, G.; Hoeben, F. J.; Tomovic, Z.; Lazzaroni, R.; Beljonne, D.; Meijer, E. W.; Schenning, A. P.; De Feyter, S. Surface-Induced Diastereomeric Complex Formation of a Nucleoside at the Liquid/ Solid Interface: Stereoselective Recognition and Preferential Adsorption. J. Am. Chem. Soc. 2013, 135, 9811−9819. (153) Guo, Z.; De Cat, I.; Van Averbeke, B.; Lin, J.; Wang, G.; Xu, H.; Lazzaroni, R.; Beljonne, D.; Schenning, A. P.; De Feyter, S. Affecting Surface Chirality via Multicomponent Adsorption of Chiral and Achiral Molecules. Chem. Commun. 2014, 50, 11903−11906. (154) Marinelli, F.; Sorrenti, A.; Corvaglia, V.; Leone, V.; Mancini, G. Molecular Description of the Propagation of Chirality from Molecules to Complex Systems: Different Mechanisms Controlled by Hydrophobic Interactions. Chem.Eur. J. 2012, 18, 14680−14688. (155) Watanabe, K.; Iida, H.; Akagi, K. Circularly Polarized Blue Luminescent Spherulites Consisting of Hierarchically Assembled Ionic Conjugated Polymers with a Helically π-Stacked Structure. Adv. Mater. 2012, 24, 6451−6456. (156) Kumar, M.; George, S. J. Homotropic And Heterotropic Allosteric Regulation of Supramolecular Chirality. Chem. Sci. 2014, 5, 3025−3030. (157) Escárcega-Bobadilla, M. V.; Salassa, G.; Belmonte, M. M.; Escudero-Adán, E. C.; Kleij, A. W. Versatile Switching in Substrate Topicity: Supramolecular Chirality Induction in Di- and Trinuclear Host Complexes. Chem.Eur. J. 2012, 18, 6805−6810. (158) Jin, Q.; Zhang, L.; Zhu, X.; Duan, P.; Liu, M. Amphiphilic Schiff Base Organogels: Metal-Ion-Mediated Chiral Twists and Chiral Recognition. Chem.Eur. J. 2012, 18, 4916−4922. (159) Rivera, J. M. Chiral Spaces: Dissymmetric Capsules through SelfAssembly. Science 1998, 279, 1021−1023. (160) Rivera, J. M.; Craig, S. L.; Martín, T.; Rebek, J. Chiral Guests and Their Ghosts in Reversibly Assembled Hosts. Angew. Chem., Int. Ed. 2000, 39, 2130−2132. (161) Tsunoda, Y.; Fukuta, K.; Imamura, T.; Sekiya, R.; Furuyama, T.; Kobayashi, N.; Haino, T. High Diastereoselection of a Dissymmetric Capsule by Chiral Guest Complexation. Angew. Chem., Int. Ed. 2014, 53, 7243−7247.

(162) Sobczuk, A. A.; Tsuchiya, Y.; Shiraki, T.; Tamaru, S.-i.; Shinkai, S. Creation of Chiral Thixotropic Gels through a Crown-Ammonium Interaction and their Application to a Memory-Erasing Recycle System. Chem.Eur. J. 2012, 18, 2832−2838. (163) Lv, K.; Qin, L.; Wang, X.; Zhang, L.; Liu, M. A Chiroptical Switch Based on Supramolecular Chirality Transfer Through Alkyl Chain Entanglement and Dynamic Covalent Bonding. Phys. Chem. Chem. Phys. 2013, 15, 20197−20202. (164) Molla, M. R.; Das, A.; Ghosh, S. Chiral Induction by Helical Neighbour: Spectroscopic Visualization of Macroscopic-Interaction Among Self-Sorted Donor and Acceptor Pi-Stacks. Chem. Commun. 2011, 47, 8934−8936. (165) Bosnich, B. Asymmetric Syntheses Asymmetric Transformations and Asymmetric Inductions in an Optically Active Solvent. J. Am. Chem. Soc. 1967, 89, 6143−6148. (166) Green, M. M.; Khatri, C.; Peterson, N. C. A Macromolecular Conformational Change Driven by a Minute Chiral Solvation Energy. J. Am. Chem. Soc. 1993, 115, 4941−4942. (167) Khatri, C. A.; Pavlova, Y.; Green, M. M.; Morawetz, H. Chiral Solvation As a Means to Quantitatively Characterize Preferential Solvation of a Helical Polymer in Mixed Solvents. J. Am. Chem. Soc. 1997, 119, 6991−6995. (168) Nakashima, H.; Koe, J. R.; Torimitsu, K.; Fujiki, M. Transfer and Amplification of Chiral Molecular Information to Polysilylene Aggregates. J. Am. Chem. Soc. 2001, 123, 4847−4848. (169) Stepanenko, V.; Li, X.; Gershberg, J.; Würthner, F. Evidence for Kinetic Nucleation in Helical Nanofiber Formation Directed by Chiral Solvent for a Perylene Bisimide Organogelator. Chem.Eur. J. 2013, 19, 4176−4183. (170) Lee, D.; Jin, Y.-J.; Kim, H.; Suzuki, N.; Fujiki, M.; Sakaguchi, T.; Kim, S. K.; Lee, W. E.; Kwak, G. Solvent-to-Polymer Chirality Transfer in Intramolecular Stack Structure. Macromolecules 2012, 45, 5379− 5386. (171) Nakano, Y.; Ichiyanagi, F.; Naito, M.; Yang, Y.; Fujiki, M. Chiroptical Generation and Inversion During the Mirror-SymmetryBreaking Aggregation of Dialkylpolysilanes Due to Limonene Chirality. Chem. Commun. 2012, 48, 6636−6638. (172) Zhang, W.; Yoshida, K.; Fujiki, M.; Zhu, X. Unpolarized-LightDriven Amplified Chiroptical Modulation Between Chiral Aggregation and Achiral Disaggregation of an Azobenzene-alt-Fluorene Copolymer in Limonene. Macromolecules 2011, 44, 5105−5111. (173) Isare, B.; Linares, M.; Zargarian, L.; Fermandjian, S.; Miura, M.; Motohashi, S.; Vanthuyne, N.; Lazzaroni, R.; Bouteiller, L. Chirality in Dynamic Supramolecular Nanotubes Induced by a Chiral Solvent. Chem.Eur. J. 2010, 16, 173−177. (174) Yashima, E.; Maeda, K.; Iida, H.; Furusho, Y.; Nagai, K. Helical Polymers: Synthesis, Structures, and Functions. Chem. Rev. 2009, 109, 6102−6211. (175) Yashima, E.; Matsushima, T.; Okamoto, Y. Poly((4carboxyphenyl)acetylene) as a Probe for Chirality Assignment of Amines by Circular-Dichroism. J. Am. Chem. Soc. 1995, 117, 11596− 11597. (176) Onouchi, H.; Kashiwagi, D.; Hayashi, K.; Maeda, K.; Yashima, E. Helicity Induction on Poly(phenylacetylene)s Bearing Phosphonic Acid Pendants with Chiral Amines and Memory of the Macromolecular Helicity Assisted by Interaction with Achiral Amines in Dimethyl Sulfoxide. Macromolecules 2004, 37, 5495−5503. (177) Ashida, Y.; Sato, T.; Morino, K.; Maeda, K.; Okamoto, Y.; Yashima, E. Helical Structural Change in Poly((4-carboxyphenyl)acetylene) by Acid-Base Complexation with an Optically Active Amine. Macromolecules 2003, 36, 3345−3350. (178) Maeda, K.; Morino, K.; Okamoto, Y.; Sato, T.; Yashima, E. Mechanism of Helix Induction on a Stereoregular Poly((4carboxyphenyl)acetylene) with Chiral Amines and Memory of the Macromolecular Helicity Assisted by Interaction with Achiral Amines. J. Am. Chem. Soc. 2004, 126, 4329−4342. (179) Maeda, K.; Kamiya, N.; Yashima, E. Poly(phenylacetylene)s Bearing a Peptide Pendant: Helical Conformational Changes of the 7389

DOI: 10.1021/cr500671p Chem. Rev. 2015, 115, 7304−7397

Chemical Reviews

Review

Columnar Assemblies with Phototunable Chirality. J. Am. Chem. Soc. 2012, 22, 18025−18032. (199) Hu, Q.; Wang, Y.; Jia, J.; Wang, C. S.; Feng, L.; Dong, R.; Sun, X.; Hao, J. Photoresponsive Chiral Nanotubes of Achiral Amphiphilic Azobenzene. Soft Matter 2012, 8, 11492−11498. (200) Ruiz, U.; Pagliusi, P.; Provenzano, C.; Shibaev, V. P.; Cipparrone, G. Supramolecular Chiral Structures: Smart Polymer Organization Guided by 2D Polarization Light. Adv. Funct. Mater. 2012, 22, 2964− 2970. (201) Tejedor, R. M.; Oriol, L.; Serrano, J. L.; Partal-Ureña, F.; LópezGonzález, J. J. Photoinduced Chiral Nematic Organization in an Achiral Glassy Nematic Azopolymer. Adv. Funct. Mater. 2007, 17, 3486−3492. (202) Hayasaka, H.; Miyashita, T.; Nakayama, M.; Kuwada, K.; Akagi, K. Dynamic Photoswitching of Helical Inversion in Liquid Crystals Containing Photoresponsive Axially Chiral Dopants. J. Am. Chem. Soc. 2012, 134, 3758−3765. (203) Gopal, A.; Hifsudheen, M.; Furumi, S.; Takeuchi, M.; Ajayaghosh, A. Thermally Assisted Photonic Inversion of Supramolecular Handedness. Angew. Chem., Int. Ed. 2012, 51, 10505−10509. (204) Yagai, S.; Yamauchi, M.; Kobayashi, A.; Karatsu, T.; Kitamura, A.; Ohba, T.; Kikkawa, Y. Control Over Hierarchy Levels in the SelfAssembly of Stackable Nanotoroids. J. Am. Chem. Soc. 2012, 134, 18205−18208. (205) Qi, Z.; Wu, C.; de Molina, P. M.; Sun, H.; Schulz, A.; Griesinger, C.; Gradzielski, M.; Haag, R.; Ansorge-Schumacher, M. B.; Schalley, C. A. Fibrous Networks with Incorporated Macrocycles: A Chiral StimuliResponsive Supramolecular Supergelator and Its Application to Biocatalysis in Organic Media. Chem.Eur. J. 2013, 19, 10150−10159. (206) Liu, Y.; Wang, T.; Liu, M. Supramolecular Polymer Hydrogels from Bolaamphiphilic L-Histidine and Benzene Dicarboxylic Acids: Thixotropy and Significant Enhancement of Eu-III Fluorescence. Chem.Eur. J. 2012, 18, 14650−14659. (207) Qin, L.; Duan, P.; Xie, F.; Zhang, L.; Liu, M. A Metal Ion Triggered Shrinkable Supramolecular Hydrogel and Controlled Release by an Amphiphilic Peptide Dendron. Chem. Commun. 2013, 49, 10823− 10825. (208) Ito, H.; Tsukube, H.; Shinoda, S. A Chirality Rewriting Cycle Mediated by A Dynamic Cyclen-Calcium Complex. Chem. Commun. 2012, 48, 10954−10956. (209) Ogoshi, T.; Akutsu, T.; Yamafuji, D.; Aoki, T.; Yamagishi, T. A. Solvent- and Achiral-Guest-Triggered Chiral Inversion in a Planar Chiral Pseudo[1]catenane. Angew. Chem., Int. Ed. 2013, 52, 8111−8115. (210) Kumar, M.; Jonnalagadda, N.; George, S. J. Molecular Recognition Driven Self-Assembly and Chiral Induction in Naphthalene Diimide Amphiphiles. Chem. Commun. 2012, 48, 10948−10950. (211) Ke, D.; Tang, A.; Zhan, C.; Yao, J. Conformation-Variable PDI@ Beta-Sheet Nanohelices Show Stimulus-Responsive Supramolecular Chirality. Chem. Commun. 2013, 49, 4914−4916. (212) Maity, S.; Kumar, P.; Haldar, D. Sonication-Induced Instant Amyloid-Like Fibril Formation and Organogelation by a Tripeptide. Soft Matter 2011, 7, 5239−5245. (213) Maity, S.; Sarkar, S.; Jana, P.; Maity, S. K.; Bera, S.; Mahalingam, V.; Haldar, D. Sonication-Responsive Organogelation of a Tripodal Peptide and Optical Properties of Embedded Tm3+ Nanoclusters. Soft Matter 2012, 8, 7960−7966. (214) Yu, X.; Liu, Q.; Wu, J.; Zhang, M.; Cao, X.; Zhang, S.; Wang, Q.; Chen, L.; Yi, T. Sonication-Triggered Instantaneous Gel-to-Gel Transformation. Chem.Eur. J. 2010, 16, 9099−9106. (215) Zhang, M.; Meng, L.; Cao, X.; Jiang, M.; Yi, T. Morphological Transformation between Three-Dimensional Gel Network and Spherical Vesicles via Sonication. Soft Matter 2012, 8, 4494−4498. (216) Komiya, N.; Muraoka, T.; Iida, M.; Miyanaga, M.; Takahashi, K.; Naota, T. Ultrasound-Induced Emission Enhancement Based on Structure-Dependent Homo- and Heterochiral Aggregations of Chiral Binuclear Platinum Complexes. J. Am. Chem. Soc. 2011, 133, 16054− 16061. (217) Kumar, J.; Nakashima, T.; Kawai, T. Inversion of Supramolecular Chirality in Bichromophoric Perylene Bisimides: Influence of Temperature and Ultrasound. Langmuir 2014, 30, 6030−6037.

Polymer Backbone Stimulated by the Pendant Conformational Change. Chem.Eur. J. 2004, 10, 4000−4010. (180) Nishimura, T.; Tsuchiya, K.; Ohsawa, S.; Maeda, K.; Yashima, E.; Nakamura, Y.; Nishimura, J. Macromolecular Helicity Induction on a Poly(phenylacetylene) with C2-Symmetric Chiral 60 FullereneBisadducts. J. Am. Chem. Soc. 2004, 126, 11711−11717. (181) Miyagawa, T.; Furuko, A.; Maeda, K.; Katagiri, H.; Furusho, Y.; Yashima, E. Dual Memory of Enantiomeric Helices in A Polyacetylene Induced by a Single Enantiomer. J. Am. Chem. Soc. 2005, 127, 5018− 5019. (182) Hase, Y.; Ishikawa, M.; Muraki, R.; Maeda, K.; Yashima, E. Helicity Induction in a Poly(4-carboxyphenyl isocyanide) with Chiral Amines and Memory of the Macromolecular Helicity in Aqueous Solution. Macromolecules 2006, 39, 6003−6008. (183) Maeda, K.; Mochizuki, H.; Watanabe, M.; Yashima, E. Switching of Macromolecular Helicity of Optically Active Poly(phenylacetylene)s Bearing Cyclodextrin Pendants Induced by Various External Stimuli. J. Am. Chem. Soc. 2006, 128, 7639−7650. (184) Nonokawa, R.; Yashima, E. Helicity Induction on A Poly(phenylacetylene) Derivative Bearing Aza-18-Crown-6 Ether Pendants in Water. J. Polym. Sci., Polym. Chem. 2003, 41, 1004−1013. (185) Vandeleene, S.; Verswyvel, M.; Verbiest, T.; Koeckelberghs, G. Synthesis, Chiroptical Behavior, and Sensing of Carboxylic Acid Functionalized Poly(phenylene ethynylene-alt-bithiophene)s. Macromolecules 2010, 43, 7412−7423. (186) Inouye, M.; Waki, M.; Abe, H. Saccharide-Dependent Induction of Chiral Helicity in Achiral Synthetic Hydrogen-Bonding Oligomers. J. Am. Chem. Soc. 2004, 126, 2022−2027. (187) Korevaar, P. A.; Newcomb, C. J.; Meijer, E. W.; Stupp, S. I. Pathway Selection in Peptide Amphiphile Assembly. J. Am. Chem. Soc. 2014, 136, 8540−8543. (188) Zhang, L.; Liu, M. Supramolecular Chirality and Chiral Inversion of Tetraphenylsulfonato Porphyrin Assemblies on Optically Active Polylysine. J. Phys. Chem. B 2009, 113, 14015−14020. (189) Mao, Y.; Liu, K.; Meng, L.; Chen, L.; Chen, L.; Yi, T. Solvent Induced Helical Aggregation in the Self-Assembly of Cholesterol Tailed Platinum Complexes. Soft Matter 2014, 10, 7615−7622. (190) Huang, Y.; Hu, J.; Kuang, W.; Wei, Z.; Faul, C. F. J. Modulating Helicity through Amphiphilicity-Tuning Supramolecular Interactions for the Controlled Assembly of Perylenes. Chem. Commun. 2011, 47, 5554−5556. (191) Hu, J.; Kuang, W.; Deng, K.; Zou, W.; Huang, Y.; Wei, Z.; Faul, C. F. J. Self-Assembled Sugar-Substituted Perylene Diimide Nanostructures with Homochirality and High Gas Sensitivity. Adv. Funct. Mater. 2012, 22, 4149−4158. (192) Duan, P.; Cao, H.; Zhang, L.; Liu, M. Gelation Induced Supramolecular Chirality: Chirality Transfer, Amplification and Application. Soft Matter 2014, 10, 5428−5448. (193) Gütz, C.; Hovorka, R.; Struch, N.; Bunzen, J.; Meyer-Eppler, G.; Qu, Z.; Grimme, S.; Topíc, F.; Rissanen, K.; Cetina, M.; Engeser, M.; Lützen, A. Enantiomerically Pure Trinuclear Helicates via Diastereoselective Self-Assembly and Characterization of Their Redox Chemistry. J. Am. Chem. Soc. 2014, 136, 11830−11838. (194) Adhikari, B.; Kraatz, H. B. Redox-Triggered Changes in the SelfAssembly of a Ferrocene-Peptide Conjugate. Chem. Commun. 2014, 50, 5551−5553. (195) del Barrio, J.; Tejedor, R. M.; Chinelatto, L. S.; Sánchez, C.; Piñol, M.; Oriol, L. Photocontrol of the Supramolecular Chirality Imposed by Stereocenters in Liquid Crystalline Azodendrimers. Chem. Mater. 2010, 22, 1714−1723. (196) Takaishi, K.; Kawamoto, M.; Tsubaki, K.; Furuyama, T.; Muranaka, A.; Uchiyama, M. Helical Chirality of Azobenzenes Induced by an Intramolecular Chiral Axis and Potential as Chiroptical Switches. Chem.Eur. J. 2011, 17, 1778−1782. (197) Tejedor, R. M.; Serrano, J. L.; Oriol, L. Photocontrol of Supramolecular Architecture in Azopolymers: Achiral and Chiral Aggregation. Eur. Polym. J. 2009, 45, 2564−2571. (198) Vera, F.; Luis Serrano, J.; De Santo, M. P.; Barberi, R.; Blanca Ros, M.; Sierra, T. Insight into the Supramolecular Organization of 7390

DOI: 10.1021/cr500671p Chem. Rev. 2015, 115, 7304−7397

Chemical Reviews

Review

(238) Lauceri, R.; D’Urso, A.; Mammana, A.; Purrello, R. In Electronic and Magnetic Properties of Chiral Molecules and Supramolecular Architectures; Naaman, R., Beratan, D. N., Waldeck, D. H., Eds.; Springer: Heidelberg, 2011; Vol. 298. (239) Zhao, L.; Xiang, R.; Ma, R.; Wang, X.; An, Y.; Shi, L. Chiral Conversion and Memory of TPPS J-aggregates in Complex Micelles: PEG-b-PDMAEMA/TPPS. Langmuir 2011, 27, 11554−11559. (240) Li, A.; Zhao, L.; Hao, J.; Ma, R.; An, Y.; Shi, L. Aggregation Behavior of the Template-Removed 5,10,15,20-Tetrakis(4sulfonatophenyl)porphyrin Chiral Array Directed by Poly(ethylene Glycol)-Block-Poly(L-Lysine). Langmuir 2014, 30, 4797−4805. (241) Zeng, L.; He, Y.; Dai, Z.; Wang, J.; Cao, Q.; Zhang, Y. Chiral Induction, Memory, and Amplification in Porphyrin Homoaggregates Based on Electrostatic Interactions. ChemPhysChem 2009, 10, 954−962. (242) Wang, J.; Ding, D.; Zeng, L.; Cao, Q.; He, Y.; Zhang, H. Transformation, Memorization and Amplification of Chirality in Cationic Co(III) Complex-Porphyrin Aggregates. New J. Chem. 2010, 34, 1394−1403. (243) Helmich, F.; Smulders, M. M.; Lee, C. C.; Schenning, A. P.; Meijer, E. W. Effect of Stereogenic Centers on the Self-Sorting, Depolymerization, and Atropisomerization Kinetics of Porphyrin-Based Aggregates. J. Am. Chem. Soc. 2011, 133, 12238−12246. (244) Helmich, F.; Lee, C. C.; Schenning, A. P. H. J.; Meijer, E. W. Chiral Memory via Chiral Amplification and Selective Depolymerization of Porphyrin Aggregates. J. Am. Chem. Soc. 2010, 132, 16753−16755. (245) Zhao, J.; Ruan, Y.; Zhou, R.; Jiang, Y. Memory of Chirality in JType Aggregates of an Achiral Perylene Dianhydride Dye Created in A Chiral Asymmetric Catalytic Synthesis. Chem.Sci. 2011, 2, 937−944. (246) Zhang, W.; Jin, W.; Fukushima, T.; Ishii, N.; Aida, T. Dynamic or Nondynamic? Helical Trajectory in Hexabenzocoronene Nanotubes Biased by a Detachable Chiral Auxiliary. J. Am. Chem. Soc. 2013, 135, 114−117. (247) Castilla, A. M.; Ousaka, N.; Bilbeisi, R. A.; Valeri, E.; Ronson, T. K.; Nitschke, J. R. High-Fidelity Stereochemical Memory in a (Fe4L4)-LII Tetrahedral Capsule. J. Am. Chem. Soc. 2013, 135, 17999−18006. (248) Yamaguchi, T.; Kimura, T.; Matsuda, H.; Aida, T. Macroscopic Spinning Chirality Memorized in Spin-coated Films of Spatially Designed Dendritic Zinc Porphyrin J-Aggregates. Angew. Chem., Int. Ed. 2004, 43, 6350−6355. (249) Ribo, J. M.; Crusats, J.; Sagues, F.; Claret, J.; Rubires, R. Chiral Sign Induction by Vortices during The Formation of Mesophases in Stirred Solutions. Science 2001, 292, 2063−2066. (250) Luisi, P. L. The Emergence of Life. From Chemical Origins to Synthetic Biology; Cambridge University Press: New York, 2010. (251) Qiu, H.; Che, S. Chiral Mesoporous Silica: Chiral Construction and Imprinting via Cooperative Self-Assembly of Amphiphiles and Silica Precursors. Chem. Soc. Rev. 2011, 40, 1259−1268. (252) Miyagawa, T.; Yamamoto, M.; Muraki, R.; Onouchi, H.; Yashima, E. Supramolecular Helical Assembly of An Achiral Cyanine Dye in An Induced Helical Amphiphilic Poly(phenylacetylene) Interior in Water. J. Am. Chem. Soc. 2007, 129, 3676−3682. (253) Kim, O. K.; Je, J.; Jernigan, G.; Buckley, L.; Whitten, D. SuperHelix Formation Induced by Cyanine J-Aggregates onto Random-Coil Carboxymethyl Amylose as Template. J. Am. Chem. Soc. 2006, 128, 510−516. (254) Young, W. R.; Aviram, A.; Cox, R. J. Stilbene Derivatives. New Class of Room Temperature Nematic Liquids. J. Am. Chem. Soc. 1972, 94, 3976−3981. (255) Lubensky, T. C. New Banana Phases. Science 2000, 288, 2146− 2147. (256) Keith, C.; Reddy, R. A.; Hauser, A.; Baumeister, U.; Tschierske, C. Silicon-Containing Polyphilic Bent-Core Molecules: The Importance of Nanosegregation for The Development of Chirality and Polar Order in Liquid Crystalline Phases Formed by Achiral Molecules. J. Am. Chem. Soc. 2006, 128, 3051−3066. (257) Jeong, K. U.; Yang, D. K.; Graham, M. J.; Tu, Y.; Kuo, S. W.; Knapp, B. S.; Harris, F. W.; Cheng, S. Z. D. Construction of Chiral Propeller Architectures from Achiral Molecules. Adv. Mater. 2006, 18, 3229−3232.

(218) Kurouski, D.; Dukor, R. K.; Lu, X.; Nafie, L. A.; Lednev, I. K. Spontaneous Inter-Conversion of Insulin Fibril Chirality. Chem. Commun. 2012, 48, 2837−2839. (219) Lin, R.; Zhang, H.; Li, S.; Chen, L.; Zhang, W.; Wen, T. B.; Zhang, H.; Xia, H. pH-Switchable Inversion of the Metal-Centered Chirality of Metallabenzenes: Opposite Stereodynamics in Reactions of Ruthenabenzene with L- and D-Cysteine. Chem.Eur. J. 2011, 17, 2420−2427. (220) Kurouski, D.; Lu, X.; Popova, L.; Wan, W.; Shanmugasundaram, M.; Stubbs, G.; Dukor, R. K.; Lednev, I. K.; Nafie, L. A. Is Supramolecular Filament Chirality the Underlying Cause of Major Morphology Differences in Amyloid Fibrils? J. Am. Chem. Soc. 2014, 136, 2302−2312. (221) Lifson, S.; Andreola, C.; Peterson, N. C.; Green, M. M. Macromolecular Stereochemistry-Helical Sense Preference in OpticallyActive Polyisocyanates-Amplification of a Conformational Equilibrium Deuterium-Isotope Effect. J. Am. Chem. Soc. 1989, 111, 8850−8858. (222) Palmans, A. R.; Meijer, E. W. Amplification of Chirality in Dynamic Supramolecular Aggregates. Angew. Chem., Int. Ed. 2007, 46, 8948−8968. (223) Kulkarni, C.; Munirathinam, R.; George, S. J. Self-Assembly of Coronene Bisimides: Mechanistic Insight and Chiral Amplification. Chem.Eur. J. 2013, 19, 11270−11278. (224) Smulders, M. M. J.; Stals, P. J. M.; Mes, T.; Paffen, T. F. E.; Schenning, A. P. H. J.; Palmans, A. R. A.; Meijer, E. W. Probing the Limits of the Majority-Rules Principle in a Dynamic Supramolecular Polymer. J. Am. Chem. Soc. 2010, 132, 620−626. (225) van Gestel, J.; van der Schoot, P.; Michels, M. A. J. Amplification of Chirality in Helical Supramolecular Polymers Beyond the LongChain Limit. J. Chem. Phys. 2004, 120, 8253−8261. (226) van Gestel, J. Amplification of Chirality in Helical Supramolecular Polymers: The Majority-Rules Principle. Macromolecules 2004, 37, 3894−3898. (227) Nie, B.; Zhan, T.; Zhou, T.; Xiao, Z.; Jiang, G.; Zhao, X. SelfAssembly of Chiral Propeller-like Supermolecules with Unusual ″Sergeants-and-Soldiers″ and ″Majority-Rules″ Effects. Chem.Asian J. 2014, 9, 754−758. (228) Seki, T.; Asano, A.; Seki, S.; Kikkawa, Y.; Murayama, H.; Karatsu, T.; Kitamura, A.; Yagai, S. Rational construction of perylene bisimide columnar superstructures with a biased helical sense. Chem.Eur. J. 2011, 17, 3598−3608. (229) Aparicio, F.; Vicente, F.; Sánchez, L. Amplification of Chirality in N,N′-1,2-Ethanediylbisbenzamides: From Planar Sheets to Twisted Ribbons. Chem. Commun. 2010, 46, 8356−8358. (230) Cao, H.; Zhu, X.; Liu, M. Self-Assembly of Racemic Alanine Derivatives: Unexpected Chiral Twist and Enhanced Capacity for the Discrimination of Chiral Species. Angew. Chem., Int. Ed. 2013, 52, 4122− 4126. (231) Shimomura, K.; Ikai, T.; Kanoh, S.; Yashima, E.; Maeda, K. Stirring Competes with Chemical Induction in Chiral Selection of Soft Matter Aggregates. Nat. Chem. 2014, 6, 429−434. (232) Takashima, S.; Abe, H.; Inouye, M. Copper(II)/PhenanthrolineMediated CD-Enhancement and Chiral Memory Effect on a MetaEthynylpyridine Oligomer. Chem. Commun. 2012, 48, 3330−3332. (233) Bellacchio, E.; Lauceri, R.; Gurrieri, S.; Scolaro, L. M.; Romeo, A.; Purrello, R. Template-Imprinted Chiral Porphyrin Aggregates. J. Am. Chem. Soc. 1998, 120, 12353−12354. (234) Purrello, R.; Raudino, A.; Scolaro, L. M.; Loisi, A.; Bellacchio, E.; Lauceri, R. Ternary Porphyrin Aggregates and Their Chiral Memory. J. Phys. Chem. B 2000, 104, 10900−10908. (235) Rosaria, L.; D’Urso, A.; Mammana, A.; Purrello, R. Chiral Memory: Induction, Amplification, and Switching in Porphyrin Assemblies. Chirality 2008, 20, 411−419. (236) Mammana, A.; De Napoli, M.; Lauceri, R.; Purrello, R. Induction and Memory of Chirality in Porphyrin Hetero-Aggregates: The Role of the Central Metal Ion. Bioorg. Med. Chem. 2005, 13, 5159−5163. (237) Mammana, A.; D’Urso, A.; Lauceri, R.; Purrello, R. Switching Off and On the Supramolecular Chiral Memory in Porphyrin Assemblies. J. Am. Chem. Soc. 2007, 129, 8062−8063. 7391

DOI: 10.1021/cr500671p Chem. Rev. 2015, 115, 7304−7397

Chemical Reviews

Review

(258) Hough, L. E.; Spannuth, M.; Nakata, M.; Coleman, D. A.; Jones, C. D.; Dantlgraber, G.; Tschierske, C.; Watanabe, J.; Körblova, E.; Walba, D. M.; Maclennan, J. E.; Glaser, M. A.; Clark, N. A. Chiral Isotropic Liquids from Achiral Molecules. Science 2009, 325, 452−456. (259) Hough, L. E.; Jung, H. T.; Krüerke, D.; Heberling, M. S.; Nakata, M.; Jones, C. D.; Chen, D.; Link, D. R.; Zasadzinski, J.; Heppke, G.; Rabe, J. P.; Stocker, W.; Körblova, E.; Walba, D. M.; Glaser, M. A.; Clark, N. A. Helical Nanofilament Phases. Science 2009, 325, 456−460. (260) Otani, T.; Araoka, F.; Ishikawa, K.; Takezoe, H. Enhanced Optical Activity by Achiral Rod-like Molecules Nanosegregated in The B4 Structure of Achiral Bent-Core Molecules. J. Am. Chem. Soc. 2009, 131, 12368−12372. (261) Ueda, T.; Masuko, S.; Araoka, F.; Ishikawa, K.; Takezoe, H. A General Method for The Enantioselective Formation of Helical Nanofilaments. Angew. Chem., Int. Ed. 2013, 52, 6863−6866. (262) Zhang, C.; Diorio, N.; Lavrentovich, O. D.; Jákli, A. Helical Nanofilaments of Bent-Core Liquid Crystals with A Second Twist. Nat. Commun. 2014, 5, 3302. (263) Freudenthal, J. Mueller Matrix Imaging in Crystallography. The Chiroptics of Pattern Formation; ProQuest: Ann Arbor, 2012. (264) Elemans, J. A.; Lei, S.; De Feyter, S. Molecular and Supramolecular Networks on Surfaces: from Two-Dimensional Crystal Engineering to Reactivity. Angew. Chem., Int. Ed. 2009, 48, 7298−7332. (265) Chen, Q.; Chen, T.; Wang, D.; Liu, H.-B.; Li, Y.-L.; Wan, L.-J. Structure and Structural Transition of Chiral Domains in Oligo(pPhenylenevinylene) Assembly Investigated by Scanning Tunneling Microscopy. Proc. Natl. Acad. Sci., U.S.A. 2010, 107, 2769−2774. (266) Takeuchi, M.; Tanaka, S.; Shinkai, S. On The Influence of Porphyrin π-π Stacking on Supramolecular Chirality Created in The Porphyrin-Based Twisted Tape Structure. Chem. Commun. 2005, 5539− 5541. (267) Qi, Z.; Malo de Molina, P.; Jiang, W.; Wang, Q.; Nowosinski, K.; Schulz, A.; Gradzielski, M.; Schalley, C. A. Systems Chemistry: Logic Gates Based on The Stimuli-Responsive Gel-Sol Transition of A Crown Ether-Functionalized Bis(urea) Gelator. Chem. Sci. 2012, 3, 2073−2082. (268) Nam, S. R.; Lee, H. Y.; Hong, J.-I. Control of Macroscopic Helicity by Using The Sergeants-And-Soldiers Principle in Organogels. Chem.Eur. J. 2008, 14, 6040−6043. (269) DeRossi, U.; Dähne, S.; Meskers, S. C. J.; Dekkers, H. P. J. M. Spontaneous Formation of Chirality in J-Aggregates Showing Davydov Splitting. Angew. Chem., Int. Ed. 1996, 35, 760−763. (270) Qiu, Y.; Chen, P.; Liu, M. Evolution of Various Porphyrin Nanostructures via An Oil/Aqueous Medium: Controlled SelfAssembly, Further Organization, and Supramolecular Chirality. J. Am. Chem. Soc. 2010, 132, 9644−9652. (271) Mineo, P.; Villari, V.; Scamporrino, E.; Micali, N. Supramolecular Chirality Induced by A Weak Thermal Force. Soft Matter 2014, 10, 44−47. (272) Stals, P. J. M.; Korevaar, P. A.; Gillissen, M. A. J.; de Greef, T. F. A.; Fitie, C. F. C.; Sijbesma, R. P.; Palmans, A. R. A.; Meijer, E. W. Symmetry Breaking in The Self-Assembly of Partially Fluorinated Benzene-1,3,5-Tricarboxamides. Angew. Chem., Int. Ed. 2012, 51, 11297−11301. (273) Romeo, A.; Castriciano, M. A.; Occhiuto, I.; Zagami, R.; Pasternack, R. F.; Scolaro, L. M. Kinetic Control of Chirality in Porphyrin J-Aggregates. J. Am. Chem. Soc. 2014, 136, 40−43. (274) Wang, Y.; Zhou, D.; Li, H.; Li, R.; Zhong, Y.; Sun, X.; Sun, X. Hydrogen-Bonded Supercoil Self-Assembly from Achiral Molecular Components with Light-Driven Supramolecular Chirality. J. Mater. Chem. C 2014, 2, 6402−6409. (275) Terech, P.; Weiss, R. G. Low Molecular Mass Gelators of Organic Liquids and The Properties of Their Gels. Chem. Rev. 1997, 97, 3133−3160. (276) Hirst, A. R.; Smith, D. K. Two-Component Gel-Phase MaterialsHighly Tunable Self-Assembling Systems. Chem.Eur. J. 2005, 11, 5496−5508. (277) Babu, S. S.; Praveen, V. K.; Ajayaghosh, A. Functional π-Gelators and Their Applications. Chem. Rev. 2014, 114, 1973−2129.

(278) Estroff, L. A.; Hamilton, A. D. Water Gelation by Small Organic Molecules. Chem. Rev. 2004, 104, 1201−1218. (279) Yang, Z.; Liang, G.; Xu, B. Enzymatic Hydrogelation of Small Molecules. Acc. Chem. Res. 2008, 41, 315−326. (280) Zhang, M.; Xu, D.; Yan, X.; Chen, J.; Dong, S.; Zheng, B.; Huang, F. Self-Healing Supramolecular Gels Formed by Crown Ether Based Host-Guest Interactions. Angew. Chem., Int. Ed. 2012, 51, 7011−7015. (281) Zhao, Y.; Fan, Y.; Mu, X.; Gao, H.; Wang, J.; Zhang, J.; Yang, W.; Chi, L.; Wang, Y. Self-Assembly of Luminescent Twisted Fibers Based on Achiral Quinacridone Derivatives. Nano Res. 2009, 2, 493−499. (282) Kimura, M.; Hatanaka, T.; Nomoto, H.; Takizawa, J.; Fukawa, T.; Tatewaki, Y.; Shirai, H. Self-Assembled Helical Nanofibers Made of Achiral Molecular Disks Having Molecular Adapter. Chem. Mater. 2010, 22, 5732−5738. (283) Zhang, S.; Yang, S.; Lan, J.; Yang, S.; You, J. Helical Nonracemic Tubular Coordination Polymer Gelators from Simple Achiral Molecules. Chem. Commun. 2008, 6170−6172. (284) Shen, Z.; Wang, T.; Liu, M. Macroscopic Chirality of Supramolecular Gels Formed from Achiral Tris(Ethyl Cinnamate) Benzene-1,3,5-Tricarboxamides. Angew. Chem., Int. Ed. 2014, 53, 13424−13428. (285) Huang, X.; Li, C.; Jiang, S. G.; Wang, X. S.; Zhang, B. W.; Liu, M. H. Self-Assembled Spiral Nanoarchitecture and Supramolecular Chirality in Langmuir−Blodgett Films of An Achiral Amphiphilic Barbituric Acid. J. Am. Chem. Soc. 2004, 126, 1322−1323. (286) Guo, Z.; Yuan, J.; Cui, Y.; Chang, F.; Sun, W.; Liu, M. Supramolecular Assemblies of A Series of 2-Arylbenzimidazoles at The Air/Water Interface: in Situ Coordination, Surface Architecture and Supramolecular Chirality. Chem.Eur. J. 2005, 11, 4155−4162. (287) Guo, Z.; Jiao, T.; Liu, M. Effect of Substituent Position in Coumarin Derivatives on The Interfacial Assembly: Reversible Photodimerization and Supramolecular Chirality. Langmuir 2007, 23, 1824−1829. (288) Zhang, Y.; Chen, P.; Liu, M. Supramolecular Chirality from An Achiral Azobenzene Derivative through The Interfacial Assembly: Effect of The Configuration of Azobenzene Unit. Langmuir 2006, 22, 10246− 10250. (289) Zhang, L.; Lu, Q.; Liu, M. Fabrication of Chiral Langmuir− Schaefer Films from Achiral TPPS and Amphiphiles Through The Adsorption at The Air/Water Interface. J. Phys. Chem.B 2003, 107, 2565−2569. (290) Zhang, Y.; Chen, P.; Liu, M. A General Method for Constructing Optically Active Supramolecular Assemblies from Intrinsically Achiral Water-Insoluble Free-Base Porphyrins. Chem.Eur. J. 2008, 14, 1793− 1803. (291) Yao, P.; Qiu, Y.; Chen, P.; Ma, Y.; He, S.; Zheng, J.-Y.; Liu, M. Interfacial Molecular Assemblies of Metalloporphyrins with Two Trans or One Axial Ligands. ChemPhysChem 2010, 11, 722−729. (292) Huang, X.; Liu, M. Chirality of Photopolymerized Organized Supramolecular Polydiacetylene Films. Chem. Commun. 2003, 66−67. (293) Zou, G.; Wang, Y.; Zhang, Q.; Kohn, H.; Manaka, T.; Iwamoto, M. Molecular Structure Modulated Properties of AzobenzeneSubstituted Polydiacetylene LB Films: Chirality Formation and Thermal Stability. Polymer 2010, 51, 2229−2235. (294) Chen, P. L.; Ma, X. G.; Liu, M. H. Optically Active Phthalocyaninato-Polysiloxane Constructed from Achiral Monomers: from Noncovalent Assembly to Covalent Polymer. Macromolecules 2007, 40, 4780−4784. (295) Xu, Y. Y.; Rao, Y.; Zheng, D. S.; Guo, Y.; Liu, M. H.; Wang, H. F. Inhomogeneous and Spontaneous Formation of Chirality in The Langmuir Monolayer of Achiral Molecules at The Air/Water Interface Probed by In Situ Surface Second Harmonic Generation Linear Dichroism. J. Phys. Chem. C 2009, 113, 4088−4098. (296) Wang, T.; Liu, M. Langmuir-Schaefer Films of A Set of Achiral Amphiphilic Porphyrins: Aggregation and Supramolecular Chirality. Soft Matter 2008, 4, 775−783. (297) Lin, L.; Wang, T.; Lu, Z.; Liu, M.; Guo, Y. In Situ Measurement of The Supramolecular Chirality in The Langmuir Monolayers of 7392

DOI: 10.1021/cr500671p Chem. Rev. 2015, 115, 7304−7397

Chemical Reviews

Review

Achiral Porphyrins at The Air/Aqueous Interface by Second Harmonic Generation Linear Dichroism. J. Phys. Chem. C 2014, 118, 6726−6733. (298) Crusats, J.; El-Hachemi, Z.; Ribo, J. M. Hydrodynamic Effects on Chiral Induction. Chem. Soc. Rev. 2010, 39, 569−577. (299) Escudero, C.; Crusats, J.; Diez-Perez, I.; El-Hachemi, Z.; Ribo, J. M. Folding and Hydrodynamic Forces in J-Aggregates of 5-Phenyl10,15,20-Tris(4-Sulfophenyl)Porphyrin. Angew. Chem., Int. Ed. 2006, 45, 8032−8035. (300) Tsuda, A.; Alam, M. A.; Harada, T.; Yamaguchi, T.; Ishii, N.; Aida, T. Spectroscopic Visualization of Vortex Flows Using DyeContaining Nanofibers. Angew. Chem., Int. Ed. 2007, 46, 8198−8202. (301) D’Urso, A.; Randazzo, R.; LoFaro, L.; Purrello, R. Vortexes and Nanoscale Chirality. Angew. Chem., Int. Ed. 2010, 49, 108−112. (302) Okano, K.; Arteaga, O.; Ribo, J. M.; Yamashita, T. Emergence of Chiral Environments by Effect of Flows: The Case of An Ionic Oligomer and Congo Red Dye. Chem.Eur. J. 2011, 17, 9288−9292. (303) Okano, K.; Taguchi, M.; Fujiki, M.; Yamashita, T. Circularly Polarized Luminescence of Rhodamine B in A Supramolecular Chiral Medium Formed by A Vortex Flow. Angew. Chem., Int. Ed. 2011, 50, 12474−12477. (304) Sorrenti, A.; El-Hachemi, Z.; Arteaga, O.; Canillas, A.; Crusats, J.; Ribo, J. M. Kinetic Control of The Supramolecular Chirality of Porphyrin J-Aggregates. Chem.Eur. J. 2012, 18, 8820−8826. (305) Micali, N.; Engelkamp, H.; van Rhee, P. G.; Christianen, P. C. M.; Scolaro, L. M.; Maan, J. C. Selection of Supramolecular Chirality by Application of Rotational and Magnetic Forces. Nat. Chem. 2012, 4, 201−207. (306) Petit-Garrido, N.; Claret, J.; Ignes-Mullol, J.; Sagues, F. Stirring Competes with Chemical Induction in Chiral Selection of Soft Matter Aggregates. Nat. Commun. 2012, 3, 1001. (307) Bailey, J.; Chrysostomou, A.; Hough, J. H.; Gledhill, T. M.; McCall, A.; Clark, S.; Ménard, F.; Tamura, M. Circular Polarization in Star-Formation Regions: Implications for Biomolecular Homochirality. Science 1998, 281, 672−674. (308) Cronin, J. R.; Pizzarello, S. Enantiomeric Excesses in Meteoritic Amino Acids. Science 1997, 275, 951−955. (309) Jiang, H.; Pan, X. J.; Lei, Z. Y.; Zou, G.; Zhang, Q. J.; Wang, K. Y. Control of Supramolecular Chirality for Polydiacetylene LB Films with The Command Azobenzene Derivative Monolayer. J. Mater. Chem. 2011, 21, 4518−4522. (310) Xu, Y.; Jiang, H.; Zhang, Q.; Wang, F.; Zou, G. Helical Polydiacetylene Prepared in The Liquid Crystal Phase Using Circular Polarized Ultraviolet Light. Chem. Commun. 2014, 50, 365−367. (311) Xu, Y.; Yang, G.; Xia, H.; Zou, G.; Zhang, Q.; Gao, J. Enantioselective Synthesis of Helical Polydiacetylene by Application of Linearly Polarized Light and Magnetic Field. Nat. Commun. 2014, 5, 5050. (312) Chen, P. L.; Ma, X. G.; Hu, K. M.; Rong, Y. L.; Liu, M. H. Left or Right? The Direction of Compression-Generated Vortex-Like Flow Selects The Macroscopic Chirality of Interfacial Molecular Assemblies. Chem.Eur. J. 2011, 17, 12108−12114. (313) Safont-Sempere, M. M.; Fernández, G.; Würthner, F. SelfSorting Phenomena in Complex Supramolecular Systems. Chem. Rev. 2011, 111, 5784−5814. (314) Kamada, T.; Aratani, N.; Ikeda, T.; Shibata, N.; Higuchi, Y.; Wakamiya, A.; Yamaguchi, S.; Kim, K. S.; Yoon, Z. S.; Kim, D.; Osuka, A. High Fidelity Self-Sorting Assembling of meso-Cinchomeronimide Appended meso-meso Linked Zn(II) Diporphyrins. J. Am. Chem. Soc. 2006, 128, 7670−7678. (315) Lee, S. J.; Cho, S.-H.; Mulfort, K. L.; Tiede, D. M.; Hupp, J. T.; Nguyen, S. T. Cavity-Tailored, Self-Sorting Supramolecular Catalytic Boxes for Selective Oxidation. J. Am. Chem. Soc. 2008, 130, 16828− 16829. (316) Ito, S.; Ono, K.; Iwasawa, N. Controlled Self-Assembly of Multiple Diastereomeric Macrocyclic Boronic Esters Composed of Two Chiral Units. J. Am. Chem. Soc. 2012, 134, 13962−13965. (317) Lin, J.; Guo, Z.; Plas, J.; Amabilino, D. B.; De Feyter, S.; Schenning, A. P. H. J. Homochiral and Heterochiral Assembly

Preferences at Different Length Scales-Conglomerates and Racemates in The Same Assemblies. Chem. Commun. 2013, 49, 9320−9322. (318) Safont-Sempere, M. M.; Osswald, P.; Stolte, M.; Grune, M.; Renz, M.; Kaupp, M.; Radacki, K.; Braunschweig, H.; Würthner, F. Impact of Molecular Flexibility on Binding Strength and Self-Sorting of Chiral π-Surfaces. J. Am. Chem. Soc. 2011, 133, 9580−9591. (319) Roche, C.; Sun, H. J.; Prendergast, M. E.; Leowanawat, P.; Partridge, B. E.; Heiney, P. A.; Araoka, F.; Graf, R.; Spiess, H. W.; Zeng, X.; Ungar, G.; Percec, V. Homochiral Columns Constructed by Chiral Self-Sorting during Supramolecular Helical Organization of Hat-Shaped Molecules. J. Am. Chem. Soc. 2014, 136, 7169−7185. (320) Makarević, J.; Jokić, M.; Raza, Z.; Štefanić, Z.; Kojić-Prodić, B.; Ž inić, M. Chiral Bis(Amino Alcohol)Oxalamide GelatorsGelation Properties and Supramolecular Organization: Racemate versus Pure Enantiomer Gelation. Chem.Eur. J. 2003, 9, 5567−5580. (321) Amemiya, R.; Mizutani, M.; Yamaguchi, M. Two-Component Gel Formation by Pseudoenantiomeric Ethynylhelicene Oligomers. Angew. Chem., Int. Ed. 2010, 49, 1995−1999. (322) Nagy, K. J.; Giano, M. C.; Jin, A.; Pochan, D. J.; Schneider, J. P. Enhanced Mechanical Rigidity of Hydrogels Formed from Enantiomeric Peptide Assemblies. J. Am. Chem. Soc. 2011, 133, 14975−14977. (323) Shen, Z.; Wang, T.; Liu, M. Tuning The Gelation Ability of Racemic Mixture by Melamine: Enhanced Mechanical Rigidity and Tunable Nanoscale Chirality. Langmuir 2014, 30, 10772−10778. (324) Lee, S. J.; Lin, W. Chiral Metallocycles: Rational Synthesis and Novel Applications. Acc. Chem. Res. 2008, 41, 521−537. (325) Pescitelli, G.; Di Bari, L.; Berova, N. Application of Electronic Circular Dichroism in The Study of Supramolecular Systems. Chem. Soc. Rev. 2014, 43, 5211−5233. (326) Borovkov, V. Supramolecular Chirality in Porphyrin Chemistry. Symmetry 2014, 6, 256−294. (327) Ohta, N.; Fuyuhiro, A.; Yamanari, K. Crystal Structure and Chiral Recognition of Supramolecular Complex Composed of Tris(1,2Diaminoethane)Cobalt(III) Carbonate and alpha-Cyclodextrin, [Co(en)3]2(CO3)3·2(alpha-CDX). Chem. Commun. 2010, 46, 3535−3537. (328) Kodama, K.; Kobayashi, Y.; Saigo, K. Two-Component Supramolecular Helical Architectures: Creation of Tunable Dissymmetric Cavities for The Inclusion and Chiral Recognition of The Third Components. Chem.Eur. J. 2007, 13, 2144−2152. (329) Lemieux, R. P. Molecular Recognition in Chiral Smectic Liquid Crystals: The Effect of Core-Core Interactions and Chirality Transfer on Polar Order. Chem. Soc. Rev. 2007, 36, 2033−2045. (330) Berova, N.; Pescitelli, G.; Petrovic, A. G.; Proni, G. Probing Molecular Chirality by CD-Sensitive Dimeric Metalloporphyrin Hosts. Chem. Commun. 2009, 5958−5980. (331) Ching, H. Y.; Clifford, S.; Bhadbhade, M.; Clarke, R. J.; Rendina, L. M. Synthesis and Supramolecular Studies of Chiral Boronated Platinum(II) Complexes: Insights into The Molecular Recognition of Carboranes by beta-Cyclodextrin. Chem.Eur. J. 2012, 18, 14413− 14425. (332) Petrovic, A. G.; Chen, Y.; Pescitelli, G.; Berova, N.; Proni, G. CD-Sensitive Zn-Porphyrin Tweezer Host-Guest Complexes, part 1: MC/OPLS-2005 Computational Approach for Predicting Preferred Interporphyrin Helicity. Chirality 2010, 22, 129−139. (333) Borovkov, V. V.; Hembury, G. A.; Inoue, Y. Origin, Control, and Application of Supramolecular Chirogenesis in Bisporphyrin-Based Systems. Acc. Chem. Res. 2004, 37, 449−459. (334) Fiedler, D.; Leung, D. H.; Bergman, R. G.; Raymond, K. N. Selective Molecular Recognition, C-H Bond Activation, and Catalysis in Nanoscale Reaction Vessels. Acc. Chem. Res. 2005, 38, 349−358. (335) Jo, H. H.; Lin, C.-Y.; Anslyn, E. V. Rapid Optical Methods for Enantiomeric Excess Analysis: from Enantioselective Indicator Displacement Assays to Exciton-Coupled Circular Dichroism. Acc. Chem. Res. 2014, 47, 2212−2221. (336) Zhang, D.-W.; Zhao, X.; Li, Z.-T. Aromatic Amide and Hydrazide Foldamer-Based Responsive Host-Guest Systems. Acc. Chem. Res. 2014, 47, 1961−1970. 7393

DOI: 10.1021/cr500671p Chem. Rev. 2015, 115, 7304−7397

Chemical Reviews

Review

(357) Edwards, W.; Smith, D. K. Enantioselective Component Selection in Multicomponent Supramolecular Gels. J. Am. Chem. Soc. 2014, 136, 1116−1124. (358) Zhang, X.; Yin, J.; Yoon, J. Recent Advances in Development of Chiral Fluorescent and Colorimetric Sensors. Chem. Rev. 2014, 114, 4918−4959. (359) Qing, G.; Sun, T. The Transformation of Chiral Signals into Macroscopic Properties of Materials Using Chirality-Responsive Polymers. NPG Asia Mater. 2012, 4, e4/1. (360) Bentley, K. W.; Wolf, C. Stereodynamic Chemosensor with Selective Circular Dichroism and Fluorescence Readout for In Situ Determination of Absolute Configuration, Enantiomeric Excess, and Concentration of Chiral Compounds. J. Am. Chem. Soc. 2013, 135, 12200−12203. (361) Wolf, C.; Bentley, K. W. Chirality Sensing Using Stereodynamic Probes with Distinct Electronic Circular Dichroism Output. Chem. Soc. Rev. 2013, 42, 5408−5424. (362) Huang, X.; Borhan, B.; Rickman, B. H.; Nakanishi, K.; Berova, N. Zinc Porphyrin Tweezer in Host-Guest Complexation: Determination of Absolute Configurations of Primary Monoamines by Circular Dichroism. Chem.Eur. J. 2000, 6, 216−224. (363) Huang, X.; Fujioka, N.; Pescitelli, G.; Koehn, F. E.; Williamson, R. T.; Nakanishi, K.; Berova, N. Absolute Configurational Assignments of Secondary Amines by CD-Sensitive Dimeric Zinc Porphyrin Host. J. Am. Chem. Soc. 2002, 124, 10320−10335. (364) Huang, X.; Rickman, B. H.; Borhan, B.; Berova, N.; Nakanishi, K. Zinc Porphyrin Tweezer in Host-Guest Complexation: Determination of Absolute Configurations of Diamines, Amino Acids, and Amino Alcohols by Circular Dichroism. J. Am. Chem. Soc. 1998, 120, 6185− 6186. (365) Kurtan, T.; Nesnas, N.; Koehn, F. E.; Li, Y.-Q.; Nakanishi, K.; Berova, N. Chiral Recognition by CD-Sensitive Dimeric Zinc Porphyrin Host. 2. Structural Studies of Host−Guest Complexes with Chiral Alcohol and Monoamine Conjugates. J. Am. Chem. Soc. 2001, 123, 5974−5982. (366) Kurtan, T.; Nesnas, N.; Li, Y.-Q.; Huang, X.; Nakanishi, K.; Berova, N. Chiral Recognition by CD-Sensitive Dimeric Zinc Porphyrin Host. 1. Chiroptical Protocol for Absolute Configurational Assignments of Monoalcohols and Primary Monoamines. J. Am. Chem. Soc. 2001, 123, 5962−5973. (367) Proni, G.; Pescitelli, G.; Huang, X.; Nakanishi, K.; Berova, N. Magnesium Tetraarylporphyrin Tweezer: A CD-Sensitive Host for Absolute Configurational Assignments of α-Chiral Carboxylic Acids. J. Am. Chem. Soc. 2003, 125, 12914−12927. (368) Li, X.; Tanasova, M.; Vasileiou, C.; Borhan, B. Fluorinated Porphyrin Tweezer: A Powerful Reporter of Absolute Configuration for Erythro and Threo Diols, Amino Alcohols, and Diamines. J. Am. Chem. Soc. 2008, 130, 1885−1893. (369) Katoono, R.; Kawai, H.; Fujiwara, K.; Suzuki, T. Dynamic Molecular Propeller: Supramolecular Chirality Sensing by Enhanced Chiroptical Response through The Transmission of Point Chirality to Mobile Helicity. J. Am. Chem. Soc. 2009, 131, 16896−16904. (370) Feng, L.; Zhao, C.; Xiao, Y.; Wu, L.; Ren, J.; Qu, X. Electrochemical DNA Three-Way Junction Based Sensor for Distinguishing Chiral Metallo-Supramolecular Complexes. Chem. Commun. 2012, 48, 6900−6902. (371) Feng, L.; Xu, B.; Ren, J.; Zhao, C.; Qu, X. A Human Telomeric DNA-Based Chiral Biosensor. Chem. Commun. 2012, 48, 9068−9070. (372) Aimi, J.; Oya, K.; Tsuda, A.; Aida, T. Chiroptical Sensing of Asymmetric Hydrocarbons Using A Homochiral Supramolecular Box from a Bismetalloporphyrin Rotamer. Angew. Chem., Int. Ed 2007, 46, 2031−2035. (373) You, L.; Berman, J. S.; Anslyn, E. V. Dynamic Multi-Component Covalent Assembly for The Reversible Binding of Secondary Alcohols and Chirality Sensing. Nat. Chem. 2011, 3, 943−948. (374) Biedermann, F.; Nau, W. M. Noncovalent Chirality Sensing Ensembles for The Detection and Reaction Monitoring of Amino Acids, Peptides, Proteins, and Aromatic Drugs. Angew. Chem., Int. Ed. 2014, 53, 5694−5699.

(337) Mutihac, L.; Lee, J. H.; Kim, J. S.; Vicens, J. Recognition of Amino Acids by Functionalized Calixarenes. Chem. Soc. Rev. 2011, 40, 2777−2779. (338) Harada, A. Cyclodextrin-Based Molecular Machines. Acc. Chem. Res. 2001, 34, 456−464. (339) Chankvetadze, B. Combined Approach Using Capillary Electrophoresis and NMR Spectroscopy for An Understanding of Enantioselective Recognition Mechanisms by Cyclodextrins. Chem. Soc. Rev. 2004, 33, 337−347. (340) Dong, S.; Zheng, B.; Wang, F.; Huang, F. Supramolecular Polymers Constructed from Macrocycle-Based Host-Guest Molecular Recognition Motifs. Acc. Chem. Res. 2014, 47, 1982−1994. (341) Ghale, G.; Nau, W. M. Dynamically Analyte-Responsive Macrocyclic Host-Fuorophore Systems. Acc. Chem. Res. 2014, 47, 2150−2159. (342) Liu, Y.; Chen, Y. Cooperative Binding and Multiple Recognition by Bridged Bis(β-Cyclodextrin)s with Functional Linkers. Acc. Chem. Res. 2006, 39, 681−691. (343) Ma, X.; Tian, H. Stimuli-Responsive Supramolecular Polymers in Aqueous Solution. Acc. Chem. Res. 2014, 47, 1971−1981. (344) Wang, M.-X. Nitrogen and Oxygen Bridged Calixaromatics: Synthesis, Structure, Functionalization, and Molecular Recognition. Acc. Chem. Res. 2012, 45, 182−195. (345) Zhang, M.; Yan, X.; Huang, F.; Niu, Z.; Gibson, H. W. StimuliResponsive Host-Guest Systems Based on The Recognition of Cryptands by Organic Guests. Acc. Chem. Res. 2014, 47, 1995−2005. (346) Bellia, F.; La Mendola, D.; Pedone, C.; Rizzarelli, E.; Saviano, M.; Vecchio, G. Selectively Functionalized Cyclodextrins and Their Metal Complexes. Chem. Soc. Rev. 2009, 38, 2756−2781. (347) Jung, J. H.; Moon, S.-J.; Ahn, J.; Jaworski, J.; Shinkai, S. Controlled Supramolecular Assembly of Helical Silica NanotubeGraphene Hybrids for Chiral Transcription and Separation. ACS Nano 2013, 7, 2595−2601. (348) So, S. M.; Moozeh, K.; Lough, A. J.; Chin, J. Highly Stereoselective Recognition and Deracemization of Amino Acids by Supramolecular Self-Assembly. Angew. Chem., Int. Ed. 2014, 53, 829− 832. (349) Kodama, K.; Kobayashi, Y.; Saigo, K. Two-Component Supramolecular Helical Architectures: Creation of Tunable Dissymmetric Cavities for The Inclusion and Chiral Recognition of The Third Components. Chem.Eur. J. 2007, 13, 2144−2152. (350) Zhang, L.; Qin, L.; Wang, X.; Cao, H.; Liu, M. Supramolecular Chirality in Self-Assembled Soft Materials: Regulation of Chiral Nanostructures and Chiral Functions. Adv. Mater. 2014, 26, 6959− 6964. (351) Rekharsky, M. V.; Yamamura, H.; Inoue, C.; Kawai, M.; Osaka, I.; Arakawa, R.; Shiba, K.; Sato, A.; Ko, Y. H.; Selvapalam, N.; Kim, K.; Inoue, Y. Chiral Recognition in Cucurbituril Cavities. J. Am. Chem. Soc. 2006, 128, 14871−14880. (352) Maeda, K.; Mochizuki, H.; Osato, K.; Yashima, E. StimuliResponsive Helical Poly(Phenylacetylene)s Bearing Cyclodextrin Pendants That Exhibit Enantioselective Gelation in Response to Chirality of A Chiral Amine and Hierarchical Super-Structured Helix Formation. Macromolecules 2011, 44, 3217−3226. (353) Yoshihara, D.; Tsuchiya, Y.; Noguchi, T.; Yamamoto, T.; Dawn, A.; Shinkai, S. Cyclodextrin-Assisted Synthesis of A Metallosupramolecular Terbium(III) Polymer and Its Fluorescence Properties and Chiral Recognition. Chem.Eur. J. 2013, 19, 15485−15488. (354) Chen, X.; Huang, Z.; Chen, S.-Y.; Li, K.; Yu, X.-Q.; Pu, L. Enantioselective Gel Collapsing: A New Means of Visual Chiral Sensing. J. Am. Chem. Soc. 2010, 132, 7297−7299. (355) Tu, T.; Fang, W.; Bao, X.; Li, X.; Dotz, K. H. Visual Chiral Recognition through Enantioselective Metallogel Collapsing: Synthesis, Characterization, and Application of Platinum-Steroid Low-MolecularMass Gelators. Angew. Chem., Int. Ed. 2011, 50, 6601−6605. (356) Jintoku, H.; Takafuji, M.; Oda, R.; Ihara, H. Enantioselective Recognition by a Highly Ordered Porphyrin-Assembly on A Chiral Molecular Gel. Chem. Commun. 2012, 48, 4881−4883. 7394

DOI: 10.1021/cr500671p Chem. Rev. 2015, 115, 7304−7397

Chemical Reviews

Review

(375) Sharma, A.; Mori, T.; Lee, H. C.; Worden, M.; Bidwell, E.; Hegmann, T. Detecting, Visualizing, and Measuring Gold Nanoparticle Chirality Using Helical Pitch Measurements in Nematic Liquid Crystal Phases. ACS Nano 2014, 8, 11966−11976. (376) Rodrigues, S. P.; Lan, S.; Kang, L.; Cui, Y.; Cai, W. Nonlinear Imaging and Spectroscopy of Chiral Metamaterials. Adv. Mater. 2014, 26, 6157−6162. (377) Li, Z.; Mutlu, M.; Ozbay, E. Chiral Metamaterials: from Optical Activity and Negative Refractive Index to Asymmetric Transmission. J. Opt. 2013, 15, 023001/1. (378) Nair, G.; Singh, H. J.; Paria, D.; Venkatapathi, M.; Ghosh, A. Plasmonic Interactions at Close Proximity in Chiral Geometries: Route toward Broadband Chiroptical Response and Giant Enantiomeric Sensitivity. J. Phys. Chem. C 2014, 118, 4991−4997. (379) Cui, Y.; Kang, L.; Lan, S.; Rodrigues, S.; Cai, W. Giant Chiral Optical Response from a Twisted-Arc Metamaterial. Nano Lett. 2014, 14, 1021−1025. (380) He, Y.; Larsen, G. K.; Ingram, W.; Zhao, Y. Tunable ThreeDimensional Helically Stacked Plasmonic Layers on Nanosphere Monolayers. Nano Lett. 2014, 14, 1976−1981. (381) Hendry, E.; Carpy, T.; Johnston, J.; Popland, M.; Mikhaylovskiy, R. V.; Lapthorn, A. J.; Kelly, S. M.; Barron, L. D.; Gadegaard, N.; Kadodwala, M. Ultrasensitive Detection and Characterization of Biomolecules Using Superchiral Fields. Nat. Nanotechnol. 2010, 5, 783−787. (382) Qiu, Y.; Chen, P.; Guo, P.; Li, Y.; Liu, M. Supramolecular Chiroptical Switches Based on Achiral Molecules. Adv. Mater. 2008, 20, 2908−2913. (383) Hecht, S. Optical Switching of Hierarchical Self-Assembly: Towards “Enlightened” Materials. Small 2005, 1, 26−29. (384) Browne, W. R.; Feringa, B. L. Chiroptical Molecular Switches. In Molecular Switches, 2nd completely revised and enlarged ed.; WileyVCH Verlag GmbH & Co. KGaA: Weinheim, 2011; pp 121−179. (385) Pijper, D.; Jongejan, M. G. M.; Meetsma, A.; Feringa, B. L. LightControlled Supramolecular Helicity of a Liquid Crystalline Phase Using a Helical Polymer Functionalized with a Single Chiroptical Molecular Switch. J. Am. Chem. Soc. 2008, 130, 4541−4552. (386) Suk, J.-m.; Naidu, V. R.; Liu, X.; Lah, M. S.; Jeong, K.-S. A Foldamer-Based Chiroptical Molecular Switch That Displays Complete Inversion of The Helical Sense upon Anion Binding. J. Am. Chem. Soc. 2011, 133, 13938−13941. (387) Guo, P.; Zhang, L.; Liu, M. A Supramolecular Chiroptical Switch Exclusively from an Achiral Amphiphile. Adv. Mater. 2006, 18, 177−180. (388) Zou, G.; Jiang, H.; Zhang, Q.; Kohn, H.; Manaka, T.; Iwamoto, M. Chiroptical Switch Based on Azobenzene-Substituted Polydiacetylene LB Films under Thermal and Photic Stimuli. J. Mater. Chem. 2010, 20, 285−291. (389) Duan, P.; Qin, L.; Liu, M. Langmuir-Blodgett Films and Chiroptical Switch of an Azobenzene-Containing Dendron Regulated by The In Situ Host-Guest Reaction at The Air/Water Interface. Langmuir 2011, 27, 1326−1331. (390) Yan, N.; He, G.; Zhang, H.; Ding, L.; Fang, Y. Glucose-Based Fluorescent Low-Molecular Mass Compounds: Creation of Simple and Versatile Supramolecular Gelators. Langmuir 2010, 26, 5909−5917. (391) Kim, M.-J.; Yoo, S.-J.; Kim, D.-Y. A Supramolecular Chiroptical Switch Using an Amorphous Azobenzene Polymer. Adv. Funct. Mater. 2006, 16, 2089−2094. (392) Canary, J. W. Redox-Triggered Chiroptical Molecular Switches. Chem. Soc. Rev. 2009, 38, 747−756. (393) Tietze, L. F.; Dufert, A.; Lotz, F.; Soelter, L.; Oum, K.; Lenzer, T.; Beck, T.; Herbst-Irmer, R. Synthesis of Chiroptical Molecular Switches by Pd-Catalyzed Domino Reactions. J. Am. Chem. Soc. 2009, 131, 17879−17884. (394) Anger, E.; Srebro, M.; Vanthuyne, N.; Toupet, L.; Rigaut, S.; Roussel, C.; Autschbach, J.; Crassous, J.; Reau, R. Multifunctional and Reactive Enantiopure Organometallic Helicenes: Tuning Chiroptical Properties by Structural Variations of Mono- and Bis(Platinahelicene)s. J. Am. Chem. Soc. 2012, 134, 15628−15631.

(395) Wang, X.; Bergenfeld, I.; Arora, P. S.; Canary, J. W. Reversible Redox Reconfiguration of Secondary Structures in a Designed Peptide. Angew. Chem., Int. Ed. 2012, 51, 12099−12101. (396) Pospisil, L.; Bednarova, L.; Stepanek, P.; Slavicek, P.; Vavra, J.; Hromadova, M.; Dlouha, H.; Tarabek, J.; Teply, F. Intense Chiroptical Switching in a Dicationic Helicene-Like Derivative: Exploration of a Viologen-Type Redox Manifold of a Non-Racemic Helquat. J. Am. Chem. Soc. 2014, 136, 10826−10829. (397) Goto, H.; Yashima, E. Electron-Induced Switching of The Supramolecular Chirality of Optically Active Polythiophene Aggregates. J. Am. Chem. Soc. 2002, 124, 7943−7949. (398) Keith, C.; Amaranatha Reddy, R.; Baumeister, U.; Tschierske, C. Banana-Shaped Liquid Crystals with Two Oligosiloxane End-Groups: Field-Induced Switching of Supramolecular Chirality. J. Am. Chem. Soc. 2004, 126, 14312−14313. (399) Ter Wiel, M. K. J.; Van Delden, R. A.; Meetsma, A.; Feringa, B. L. Increased Speed of Rotation for The Smallest Light-Driven Molecular Motor. J. Am. Chem. Soc. 2003, 125, 15076−15086. (400) Pijper, D.; Feringa, B. L. Molecular Transmission: Controlling The Twist Sense of a Helical Polymer with a Single Light-Driven Molecular Motor. Angew. Chem., Int. Ed. 2007, 46, 3693−3696. (401) Wang, J.; Feringa, B. L. Dynamic Control of Chiral Space in a Catalytic Asymmetric Reaction Using a Molecular Motor. Science 2011, 331, 1429−1432. (402) Barbera, J.; Giorgini, L.; Paris, F.; Salatelli, E.; Tejedor, R. M.; Angiolini, L. Supramolecular Chirality and Reversible Chiroptical Switching in New Chiral Liquid-Crystal Azopolymers. Chem.Eur. J. 2008, 14, 11209−11221. (403) Gao, C.; Silvi, S.; Ma, X.; Tian, H.; Credi, A.; Venturi, M. Chiral Supramolecular Switches Based on (R)-Binaphthalene-Bipyridinium Guests and Cucurbituril Hosts. Chem.Eur. J. 2012, 18, 16911−16921. (404) Cao, H.; Jiang, J.; Zhu, X.; Duan, P.; Liu, M. Hierarchical Coassembly of Chiral Lipid Nanotubes with an Azobenzene Derivative: Optical and Chiroptical Switching. Soft Matter 2011, 7, 4654−4660. (405) Cesar, V.; Bellemin-Laponnaz, S.; Gade, L. H. Chiral NHeterocyclic Carbenes as Stereodirecting Ligands in Asymmetric Catalysis. Chem. Soc. Rev. 2004, 33, 619−636. (406) Ding, K.; Wang, Z.; Wang, X.; Liang, Y.; Wang, X. SelfSupported Chiral Catalysts for Heterogeneous Enantioselective Reactions. Chem.Eur. J. 2006, 12, 5188−5197. (407) Li, C.; Zhang, H.; Jiang, D.; Yang, Q. Chiral Catalysis in Nanopores of Mesoporous Materials. Chem. Commun. 2007, 547−548. (408) Luo, S.; Zhang, L.; Cheng, J.-P. Functionalized Chiral Ionic Liquids: a New Type of Asymmetric Organocatalysts and Nonclassical Chiral Ligands. Chem. Asian J. 2009, 4, 1184−1195. (409) Wohlgemuth, R. Asymmetric Biocatalysis with Microbial Enzymes and Cells. Curr. Opin. Microbiol. 2010, 13, 283−292. (410) Xie, J.-H.; Zhu, S.-F.; Zhou, Q.-L. Recent Advances in Transition Metal-Catalyzed Enantioselective Hydrogenation of Unprotected Enamines. Chem. Soc. Rev. 2012, 41, 4126−4139. (411) Lee, Y. S.; Alam, M. M.; Keri, R. S. Enantioselective Reactions of N-Acyliminium Ions Using Chiral Organocatalysts. Chem. Asian. J. 2013, 8, 2906−2919. (412) Ding, K. Synergistic Effect of Binary Component Ligands in Chiral Catalyst Library Engineering for Enantioselective Reactions. Chem. Commun. 2008, 909−921. (413) Hoveyda, A. H.; Malcolmson, S. J.; Meek, S. J.; Zhugralin, A. R. Catalytic Enantioselective Olefin Metathesis in Natural Product Synthesis. Chiral Metal-Based Complexes That Deliver High Enantioselectivity and More. Angew. Chem., Int. Ed. 2010, 49, 34−44. (414) Palomo, C.; Oiarbide, M.; Garcia, J. M. Current Progress in The Asymmetric Aldol Addition Reaction. Chem. Soc. Rev. 2004, 33, 65−75. (415) Ouellet, S. G.; Walji, A. M.; MacMillan, D. W. C. Enantioselective Organocatalytic Transfer Hydrogenation Reactions Using Hantzsch Esters. Acc. Chem. Res. 2007, 40, 1327−1339. (416) Mlynarski, J.; Paradowska, J. Catalytic Asymmetric Aldol Reactions in Aqueous Media. Chem. Soc. Rev. 2008, 37, 1502−1511. 7395

DOI: 10.1021/cr500671p Chem. Rev. 2015, 115, 7304−7397

Chemical Reviews

Review

(417) Yoon, M.; Srirambalaji, R.; Kim, K. Homochiral Metal-Organic Frameworks for Asymmetric Heterogeneous Catalysis. Chem. Rev. 2011, 112, 1196−1231. (418) Ke, C.; Yang, C.; Mori, T.; Wada, T.; Liu, Y.; Inoue, Y. Catalytic Enantiodifferentiating Photocyclodimerization of 2-Anthracenecarboxylic Acid Mediated by a Non-Sensitizing Chiral Metallosupramolecular Host. Angew. Chem., Int. Ed. 2009, 48, 6675−6677. (419) Wang, Y.; Chen, H.; Xiao, Y.; Ng, C. H.; Oh, T. S.; Tan, T. T. Y.; Ng, S. C. Preparation of Cyclodextrin Chiral Stationary Phases by Organic Soluble Catalytic ’Click’ Chemistry. Nat. Protoc. 2011, 6, 935− 942. (420) Hu, S.; Li, J.; Xiang, J.; Pan, J.; Luo, S.; Cheng, J.-P. Asymmetric Supramolecular Primary Amine Catalysis in Aqueous Buffer: Connections of Selective Recognition and Asymmetric Catalysis. J. Am. Chem. Soc. 2010, 132, 7216−7228. (421) Yang, C.; Inoue, Y. Supramolecular Photochirogenesis. In Supramolecular Photochemistry; John Wiley & Sons, Inc.: New York, 2011; pp 115−153. (422) Yang, C.; Inoue, Y. Supramolecular Photochirogenesis. Chem. Soc. Rev. 2014, 43, 4123−4143. (423) Shi, L.; Wang, X.; Sandoval, C. A.; Li, M.; Qi, Q.; Li, Z.; Ding, K. Engineering a Polymeric Chiral Catalyst by Using Hydrogen Bonding and Coordination Interactions. Angew. Chem., Int. Ed. 2006, 45, 4108− 4112. (424) Yu, L.; Wang, Z.; Wu, J.; Tu, S.; Ding, K. Directed Orthogonal Self-Assembly of Homochiral Coordination Polymers for Heterogeneous Enantioselective Hydrogenation. Angew. Chem., Int. Ed. 2010, 49, 3627−3630. (425) Uraguchi, D.; Ueki, Y.; Ooi, T. Chiral Organic Ion Pair Catalysts Assembled through a Hydrogen-Bonding Network. Science 2009, 326, 120−123. (426) Hatano, M.; Mizuno, T.; Izumiseki, A.; Usami, R.; Asai, T.; Akakura, M.; Ishihara, K. Enantioselective Diels-Alder Reactions with Anomalous Endo/Exo Selectivities Using Conformationally Flexible Chiral Supramolecular Catalysts. Angew. Chem., Int. Ed. 2011, 50, 12189−12192. (427) Xia, A. B.; Xu, D. Q.; Wu, C.; Zhao, L.; Xu, Z. Y. Organocatalytic Diels-Alder Reactions Catalysed by Supramolecular Self-Assemblies Formed from Chiral Amines and Poly(Alkene Glycol)s. Chem.Eur. J. 2012, 18, 1055−1059. (428) Jin, Q.; Zhang, L.; Cao, H.; Wang, T.; Zhu, X.; Jiang, J.; Liu, M. Self-Assembly of Copper(II) Ion-Mediated Nanotube and Its Supramolecular Chiral Catalytic Behavior. Langmuir 2011, 27, 13847−13853. (429) Qin, L.; Zhang, L.; Jin, Q.; Zhang, J.; Han, B.; Liu, M. Supramolecular Assemblies of Amphiphilic L-Proline Regulated by Compressed CO2 as a Recyclable Organocatalyst for The Asymmetric Aldol Reaction. Angew. Chem., Int. Ed. 2013, 52, 7761−7765. (430) Terashima, T.; Mes, T.; De Greef, T. F.; Gillissen, M. A.; Besenius, P.; Palmans, A. R.; Meijer, E. W. Single-Chain Folding of Polymers for Catalytic Systems in Water. J. Am. Chem. Soc. 2011, 133, 4742−4745. (431) Iida, H.; Iwahana, S.; Mizoguchi, T.; Yashima, E. Main-Chain Optically Active Riboflavin Polymer for Asymmetric Catalysis and Its Vapochromic Behavior. J. Am. Chem. Soc. 2012, 134, 15103−15113. (432) Roelfes, G.; Feringa, B. L. DNA-Based Asymmetric Catalysis. Angew. Chem., Int. Ed. 2005, 44, 3230−3232. (433) Coquiere, D.; Feringa, B. L.; Roelfes, G. DNA-Based Catalytic Enantioselective Michael Reactions in Water. Angew. Chem., Int. Ed. 2007, 46, 9308−9311. (434) Boersma, A. J.; Feringa, B. L.; Roelfes, G. Enantioselective Friedel-Crafts Reactions in Water Using a DNA-Based Catalyst. Angew. Chem., Int. Ed. 2009, 48, 3346−3348. (435) Oelerich, J.; Roelfes, G. DNA-Based Asymmetric Organometallic Catalysis in Water. Chem. Sci. 2013, 4, 2013−2017. (436) Boersma, A. J.; Coquière, D.; Geerdink, D.; Rosati, F.; Feringa, B. L.; Roelfes, G. Catalytic Enantioselective Syn Hydration of Enones in Water Using a DNA-Based Catalyst. Nat. Chem. 2010, 2, 991−995.

(437) Roe, S.; Ritson, D. J.; Garner, T.; Searle, M.; Moses, J. E. Tuneable DNA-Based Asymmetric Catalysis Using a G-Quadruplex Supramolecular Assembly. Chem. Commun. 2010, 46, 4309−4311. (438) Wang, C.; Jia, G.; Zhou, J.; Li, Y.; Liu, Y.; Lu, S.; Li, C. Enantioselective Diels−Alder Reactions with G-Quadruplex DNABased Catalysts. Angew. Chem., Int. Ed. 2012, 51, 9352−9355. (439) Wang, J.; Benedetti, E.; Bethge, L.; Vonhoff, S.; Klussmann, S.; Vasseur, J.-J.; Cossy, J.; Smietana, M.; Arseniyadis, S. DNA vs. MirrorImage DNA: a Universal Approach to Tune The Absolute Configuration in DNA-Based Asymmetric Catalysis. Angew. Chem., Int. Ed. 2013, 52, 11546−11549. (440) Pace, T. C. S.; Müller, V.; Li, S.; Lincoln, P.; Andréasson, J. Enantioselective Cyclization of Photochromic Dithienylethenes Bound to DNA. Angew. Chem., Int. Ed. 2013, 52, 4393−4396. (441) Coquière, D.; Bos, J.; Beld, J.; Roelfes, G. Enantioselective Artificial Metalloenzymes Based on a Bovine Pancreatic Polypeptide Scaffold. Angew. Chem., Int. Ed. 2009, 48, 5159−5162. (442) Zheng, L.; Marcozzi, A.; Gerasimov, J. Y.; Herrmann, A. Conformationally Constrained Cyclic Peptides: Powerful Scaffolds for Asymmetric Catalysis. Angew. Chem., Int. Ed. 2014, 53, 7599−7603. (443) Autschbach, J.; Nitsch-Velasquez, L.; Rudolph, M. TimeDependent Density Functional Response Theory for Electronic Chiroptical Properties of Chiral Molecules. Top. Curr. Chem. 2011, 298, 1−98. (444) Ben-Moshe, V.; Beratan, D. N.; Nitzan, A.; Skourtis, S. S. Chiral Control of Current Transfer in Molecules. Top. Curr. Chem. 2011, 298, 259−278. (445) Mori, T.; Inoue, Y. Recent Theoretical and Experimental Advances in The Electronic Circular Dichroisms of Planar Chiral Cyclophanes. Top. Curr. Chem. 2011, 298, 99−128. (446) Lee, B.; Lee, S.-Y. Polarization-Dependent Plasmonic Chiral Devices in Nanoplasmonics 113−135, 2014, CRC Press. (447) Sato, H.; Sato, F.; Yamagishi, A. Rewritable Optical Memory in Liquid Crystals Containing Photo-Epimerizing Cr(III) Complexes. Chem. Commun. 2013, 49, 4773−4775. (448) de Vega, L.; van Cleuvenbergen, S.; Depotter, G.; Garcia-Frutos, E. M.; Gomez-Lor, B.; Omenat, A.; Tejedor, R. M.; Serrano, J. L.; Hennrich, G.; Clays, K. Nonlinear Optical Thin Film Device from a Chiral Octopolar Phenylacetylene Liquid Crystal. J. Org. Chem. 2012, 77, 10891−10896. (449) Zou, W.; Yan, Y.; Fang, J.; Yang, Y.; Liang, J.; Deng, K.; Yao, J.; Wei, Z. Biomimetic Superhelical Conducting Microfibers with Homochirality for Enantioselective Sensing. J. Am. Chem. Soc. 2013, 136, 578−581. (450) Yang, Y.; Zhang, Y.; Wei, Z. Supramolecular Helices: Chirality Transfer from Conjugated Molecules to Structures. Adv. Mater. 2013, 25, 6039−6049. (451) Verbiest, T.; Elshocht, S. V.; Kauranen, M.; Hellemans, L.; Snauwaert, J.; Nuckolls, C.; Katz, T. J.; Persoons, A. Strong Enhancement of Nonlinear Optical Properties through Supramolecular Chirality. Science 1998, 282, 913−915. (452) Réthoré, C.; Avarvari, N.; Canadell, E.; Auban-Senzier, P.; Fourmigué, M. Chiral Molecular Metals: Syntheses, Structures, and Properties of The AsF6-Salts of Racemic (±)-, (R)-, and (S)Tetrathiafulvalene−Oxazoline Derivatives. J. Am. Chem. Soc. 2005, 127, 5748−5749. (453) Yan, Y.; Wang, R.; Qiu, X.; Wei, Z. Hexagonal Superlattice of Chiral Conducting Polymers Self-Assembled by Mimicking β-Sheet Proteins with Anisotropic Electrical Transport. J. Am. Chem. Soc. 2010, 132, 12006−12012. (454) Feng, X.; Marcon, V.; Pisula, W.; Hansen, M. R.; Kirkpatrick, J.; Grozema, F.; Andrienko, D.; Kremer, K.; Müllen, K. Towards High Charge-Carrier Mobilities by Rational Design of The Shape and Periphery of Discotics. Nat. Mater. 2009, 8, 421−426. (455) Zhang, Y.; Chen, P.; Jiang, L.; Hu, W.; Liu, M. Controllable Fabrication of Supramolecular Nanocoils and Nanoribbons and Their Morphology-Dependent Photoswitching. J. Am. Chem. Soc. 2009, 131, 2756−2757. 7396

DOI: 10.1021/cr500671p Chem. Rev. 2015, 115, 7304−7397

Chemical Reviews

Review

(456) Yang, Y.; da Costa, R. C.; Fuchter, M. J.; Campbell, A. J. Circularly Polarized Light Detection by a Chiral Organic Semiconductor Transistor. Nat. Photonics 2013, 7, 634−638. (457) Mislow, K. Absolute Asymmetric Synthesis: a Commentary. Collect. Czech. Chem. Commun. 2003, 68, 849−864. (458) Carr, R.; Evans, N. H.; Parker, D. Lanthanide Complexes as Chiral Probes Exploiting Circularly Polarized Luminescence. Chem. Soc. Rev. 2012, 41, 7673−7686. (459) Bunzli, J.-C. G.; Piguet, C. Taking Advantage of Luminescent Lanthanide Ions. Chem. Soc. Rev. 2005, 34, 1048−1077. (460) Peeters, E.; Christiaans, M. P. T.; Janssen, R. A. J.; Schoo, H. F. M.; Dekkers, H. P. J. M.; Meijer, E. W. Circularly Polarized Electroluminescence from a Polymer Light-Emitting Diode. J. Am. Chem. Soc. 1997, 119, 9909−9910. (461) Geng, Y.; Trajkovska, A.; Katsis, D.; Ou, J. J.; Culligan, S. W.; Chen, S. H. Synthesis, Characterization, and Optical Properties of Monodisperse Chiral Oligofluorenes. J. Am. Chem. Soc. 2002, 124, 8337−8347. (462) Wilson, J. N.; Steffen, W.; McKenzie, T. G.; Lieser, G.; Oda, M.; Neher, D.; Bunz, U. H. F. Chiroptical Properties of Poly(pPhenyleneethynylene) Copolymers in Thin Tilms: Large g-Values. J. Am. Chem. Soc. 2002, 124, 6830−6831. (463) Satrijo, A.; Meskers, S. C. J.; Swager, T. M. Probing a Conjugated Polymer’s Transfer of Organization-Dependent Properties from Solutions to Films. J. Am. Chem. Soc. 2006, 128, 9030−9031. (464) Goto, H.; Akagi, K. Optically Active Conjugated Polymers Prepared from Achiral Monomers by Polycondensation in a Chiral Nematic Solvent. Angew. Chem., Int. Ed. 2005, 44, 4322−4328. (465) Watanabe, K.; Osaka, I.; Yorozuya, S.; Akagi, K. Helically πStacked Thiophene-Based Copolymers with Circularly Polarized Fluorescence: High Dissymmetry Factors Enhanced by Self-Ordering in Chiral Nematic Liquid Crystal Phase. Chem. Mater. 2012, 24, 1011− 1024. (466) Phillips, K. E. S.; Katz, T. J.; Jockusch, S.; Lovinger, A. J.; Turro, N. J. Synthesis and Properties of an Aggregating Heterocyclic Helicene. J. Am. Chem. Soc. 2001, 123, 11899−11907. (467) Field, J. E.; Muller, G.; Riehl, J. P.; Venkataraman, D. Circularly Polarized Luminescence from Bridged Triarylamine Helicenes. J. Am. Chem. Soc. 2003, 125, 11808−11809. (468) Maeda, H.; Bando, Y.; Shimomura, K.; Yamada, I.; Naito, M.; Nobusawa, K.; Tsumatori, H.; Kawai, T. Chemical-Stimuli-Controllable Circularly Polarized Luminescence from Anion-Responsive π-Conjugated Molecules. J. Am. Chem. Soc. 2011, 133, 9266−9269. (469) Haketa, Y.; Bando, Y.; Takaishi, K.; Uchiyama, M.; Muranaka, A.; Naito, M.; Shibaguchi, H.; Kawai, T.; Maeda, H. Asymmetric Induction in The Preparation of Helical Receptor−Anion Complexes: Ion-Pair Formation with Chiral Cations. Angew. Chem., Int. Ed. 2012, 51, 7967− 7971. (470) Kumar, J.; Nakashima, T.; Tsumatori, H.; Mori, M.; Naito, M.; Kawai, T. Circularly Polarized Luminescence in Supramolecular Assemblies of Chiral Bichromophoric Perylene Bisimides. Chem. Eur. J. 2013, 19, 14090−14097. (471) Ikeda, M.; Tanida, T.; Yoshii, T.; Kurotani, K.; Onogi, S.; Urayama, K.; Hamachi, I. Installing Logic-Gate Responses to a Variety of Biological Substances in Supramolecular Hydrogel−Enzyme Hybrids. Nat. Chem. 2014, 6, 511−518. (472) Li, J.; Gao, Y.; Kuang, Y.; Shi, J.; Du, X.; Zhou, J.; Wang, H.; Yang, Z.; Xu, B. Dephosphorylation of D-Peptide Derivatives to Form Biofunctional, Supramolecular Nanofibers/Hydrogels and Their Potential Applications for Intracellular Imaging and Intratumoral Chemotherapy. J. Am. Chem. Soc. 2013, 135, 9907−9914. (473) McClendon, M. T.; Stupp, S. I. Tubular Hydrogels of Circumferentially Aligned Nanofibers to Encapsulate and Orient Vascular Cells. Biomaterials 2012, 33, 5713−5722. (474) Liang, G.; Yang, Z.; Zhang, R.; Li, L.; Fan, Y.; Kuang, Y.; Gao, Y.; Wang, T.; Lu, W. W.; Xu, B. Supramolecular Hydrogel of a D-Amino Acid Dipeptide for Controlled Drug Release in Vivo. Langmuir 2009, 25, 8419−8422.

(475) Marchesan, S.; Easton, C. D.; Styan, K. E.; Waddington, L. J.; Kushkaki, F.; Goodall, L.; McLean, K. M.; Forsythe, J. S.; Hartley, P. G. Chirality Effects at Each Amino Acid Position on Tripeptide SelfAssembly into Hydrogel Biomaterials. Nanoscale 2014, 6, 5172−5180. (476) Liu, G.-F.; Zhang, D.; Feng, C.-L. Control of Three-Dimensional Cell Adhesion by The Chirality of Nanofibers in Hydrogels. Angew. Chem., Int. Ed. 2014, 53, 7789−7793. (477) Das, R. K.; Zouani, O. F.; Labrugère, C.; Oda, R.; Durrieu, M.-C. Influence of Nanohelical Shape and Periodicity on Stem Cell Fate. ACS Nano 2013, 7, 3351−3361. (478) Lv, K.; Zhang, L.; Lu, W.; Liu, M. Control of Supramolecular Chirality of Nanofibers and Its Effect on Protein Adhesion. ACS Appl. Mater. Interfaces 2014, 6, 18878−18884.

7397

DOI: 10.1021/cr500671p Chem. Rev. 2015, 115, 7304−7397