Anda di halaman 1dari 94

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, Peoples Republic of China
4.3.6. Sonication
4.3.7. pH Value
4.4. Chiral Amplication in Supramolecular Systems
4.4.1. Analogue-Induced Chiral Amplication
4.4.2. Chiral Amplication in Binary Systems
4.4.3. Chiral Amplication to Nanoscale
4.4.4. Unexpected Amplication 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. Conguration 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. -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 Eect 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, 73047397

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 lms such as Langmuir
and LangmuirBlodgett lms, and others. Although there are
several reviews detailing supramolecular chirality, self-assembly,
as well as chiral nanomaterials and nanostructures,713 the eld
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 denition, 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 amplication, 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 dier 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,1416 chirality in a
coordination system,1722 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 nds 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 dened 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
ecient 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 -helix, -sheet, and random coil structures with dierent
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
dierent types of atoms or four dierent 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, 73047397

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 Wikipedias 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 ower (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 classied
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 specic 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 dierence originates from
the dierences 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 eects 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 sergeantsoldier 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. Conguration and Conformation Chirality

Conguration refers to the permanent geometry resulting from


the spatial arrangement of a systems bonds. Conformation refers
to the spatial arrangement of substituent groups that are free to
assume dierent positions in space without breaking any bonds,
because of the freedom of bond rotation. While conguration
chirality is generally used in the case of molecular chirality, such
as the absolute conguration 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
denition.

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

xed chirality,
recognition

dynamic, sergeantsoldier 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, 73047397

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 dened 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 classied 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, 73047397

Chemical Reviews

Review

SEM, and TEM observations. Helicity can be classied 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.2527 For the readers 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
IngoldPrelog 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 and chirality terms are
used for dening coordination compounds. The enantiomers can
be designated as for a left-handed twist of the propeller
described by the ligands and 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 samples 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 at 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 ows 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 at
conductive substrate such as HOPG. The STM technique
provides a molecular level resolution and is used to directly
discriminate the absolute conguration 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 reected electrons. SEM is critical in all elds
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 (<200 nm).
In addition, using TEM, a diraction pattern can also be
obtained. TEM provides very high resolution (0.1 nm) and is
useful in determining the structures of nanomaterials. It is also
useful for observing chiral structures. For example, using TEM, a
self-assembled helical twist structure formed by a pyridinecontaining amphiphilic L-glutamide was clearly observed.30
Although all of the above techniques can be used to observe
chiral nanostructures, not all of the techniques are suitable for
these observations. It is important to select the characterization

3. CHARACTERIZATION OF SUPRAMOLECULAR
CHIRALITY
An important step in the research of supramolecular chirality is
the characterization of the chirality. Although there are many
ways to characterize the chiral features of a supramolecular
system, two classes of characterization of supramolecular
chirality are usually applied. One is the morphological
observation by various microscopes, with which one can directly
observe the chiral molecules and chiral structures. With the rapid
development of STM, AFM, SEM, and TEM technologies, direct
observation of chiral structures has been made possible, and
these techniques have signicantly assisted in the development of
research on chirality, especially in the self-assembled supramolecular systems. The other class of characterization is
spectroscopy techniques such as circular dichroism (CD),
vibrational CD (VCD), and Raman optical activity (ROA)
spectroscopy. With these techniques, the dynamic features of the
supramolecular chirality can be followed and the self-assembly
process can be unveiled. Although X-ray structural analysis is
useful in determining the absolute conguration of the chiral
molecules, it requires that samples form into crystals. Selfassembly systems are generally not crystallized and thus not
applicable for X-ray crystallography. However, the molecular
conguration and packing information from the X-ray studies
can help in understanding the self-assembly process. Herein, we
do not attempt to explain all of the possible characterization
methods in detail but provide a general introduction as to how
these techniques are applied to the study of supramolecular
chirality.
3.1. Morphology Observation

Seeing is believing. The rapid development of research in


supramolecular chiral systems is largely dependent on those
techniques that lead to direct visualization of the chiral
nanostructures. Figure 3 shows typical chiral images obtained
by these techniques and a comparison between these techniques.
7308

DOI: 10.1021/cr500671p
Chem. Rev. 2015, 115, 73047397

Chemical Reviews

Review

method based on the samples and the self-assembled systems.


Taking self-assembled gel systems as an example, AFM is
generally applicable for investigating transparent gels, while SEM
is preferable for observation of white gels, since the sizes of the
nanostructures are dierent. For use of STM, a conductive
sample is generally needed, while for TEM observation, materials
with heavy metals or metal ions are relatively easily observed.

be very useful in monitoring dynamic processes if the system


contains chiral elements.
As a result, CD spectroscopy has developed rapidly because of
its application in supramolecular systems and been extended to
investigating nanosystems such as plasmonic chirality research.
This has become the most important technique for characterizing molecular, supramolecular, and nanoassembly systems.
With respect to CD measurements of supramolecular systems,
there exist excellent reviews on this technique.34,37 Thus, herein,
only a few important measurement techniques will be
mentioned, together with an explanation of the use of CD
spectroscopy in the measurement of supramolecular chirality.
3.2.2. Measurement Aspects. The CD technique is
typically suitable for viewing samples in isotropic solutions.
Appropriate selection of the solvent, solute concentration, and
measurement cell can facilitate measurement of the CD
spectrum of a system. However, when measuring a supramolecular system such as solid or colloid dispersions, membranes
or lms, gels, or liquid crystals, many factors such as the turbidity,
birefringence, and anisotropic chiral nanostructures will seriously
aect the data collection and subsequent data interpretation. A
major interference can result from linear dichroism (LD). A
useful method for the CD measurement of lms and to exclude
the eect of LD has been proposed by Spitz et al. to distinguish
intrinsic chirality from possible parasitic artifacts, and this
method was applied to LangmuirBlodgett (LB) lms.38,39 In
some self-assembly systems, the LD could be enlarged if there is
an ordered arrangement of the molecules. Thus, simultaneous
measurement of the LD spectra may give clear warnings about
the accuracy of the CD spectra. In addition, the contribution
from linear and circular birefringence (CB) cannot be neglected
in some anisotropic liquid systems. A more powerful transmission two-modulator generalized ellipsometry is proposed to
completely determine the linear and circular birefringence (CB),
the linear dichroism and circular dichroism, and the depolarization of the sample.40 This measurement is particularly important
in symmetry-breaking systems where achiral molecules or
building blocks produce macroscopic chirality. Although
relatively little attention has been paid to these measurements
in the past, more recent literature provides a deeper understanding of these eects.
3.2.3. CD Spectra and Interpretation. In CD spectral
measurement, two types of spectra are generally obtainable, as
illustrated in Figure 4A.

3.2. Spectroscopic Methods for Characterization of Chirality

Spectroscopy provides a powerful method for detecting the


chiral characteristics of supramolecular systems. Generally,
spectroscopy methods used for characterization of molecular
chirality in solution can also be used for characterizing
supramolecular systems. Depending on the light source,
measured physical variables, and principles, the measurement
methods can be divided into linear and nonlinear optical
methods. Linear optical methods include the CD, VCD, and
VOA spectroscopies apparatus, which are commercially
available. Nonlinear optical methods include SHG and SFG
methods, which are not available as commercial instruments.
Many of these techniques have been introduced in detail in books
or reviews,3336 so they will not be described here in detail.
Among these techniques, CD spectroscopy is the most widely
used for chirality characterization. We will take CD measurement
as an example to show how supramolecular chirality is most often
characterized.
3.2.1. CD Spectra of Supramolecular Systems. Circular
dichroism (CD) is the dierential absorption of left versus right
circularly polarized light. A CD spectrum records the circular
dichroism as a function of wavelength. Thus, a CD spectrum is
strongly related to the absorption spectrum of certain
compounds. If the molecules do not contain any chromophore
or the absorption is outside the wavelength region then no CD
signal can be detected. For this reason, the two types of spectra,
UVvis and CD spectroscopies, should be considered together
in most cases.
Circular dichroism spectroscopy was developed to study
molecular chirality. However, through a series of detailed
investigations using CD spectroscopy in supramolecular systems,
it has been found that this technique is particularly useful for
monitoring self-assembled systems for reasons described below.
First, self-assembly is usually a dynamic process where
assembly and disassembly occurs simultaneously. These
dynamics generally occur in the time scale of CD measurements.
Thus, CD spectroscopy is very useful in obtaining information,
such as the formation dynamics of the chiral nanostructures, as
well as the interaction between the chiral species and the chiral
substrates.
Second, CD signals originate from the electronic transitions of
the chromophore and are generally strong and sensitive to the
chromophores as well as the chromophore packing. During selfassembly, molecular packing plays an important role in
determining the chiral nanostructures, and the CD spectrum
can give detailed information on packing. In many cases, we can
detect exciton coupling in the self-assembled systems. The
exciton CD can experimentally assign molecular conformations,
absolute congurations, and molecular interactions, even for
rather complicated supramolecular systems. In many cases, it is
more sensitive than UVvis spectra and gives more detailed
information on chiral interactions.
Third, CD spectroscopy is also useful for detecting the chiral
perturbations of the assemblies. Many self-assembled systems are
dynamic and responsive to external stimuli; thus, CD spectra will

Figure 4. (A) Typically classied CD spectra in supramolecular systems.


(B) Qualitative KramersKronig-consistent transformation of a CD
bisignate band.
7309

DOI: 10.1021/cr500671p
Chem. Rev. 2015, 115, 73047397

Chemical Reviews

Review

4.1. Supramolecular Chirality in Assemblies of Chiral


Components

Observation of a peak or valley in the CD spectrum is referred


to as the Cotton eect, which is the characteristic change in
circular dichroism in the vicinity of an absorption band of a
substance. The Cotton eect is deemed positive if the circular
dichroism rst increases as the wavelength decreases (as rst
observed by Cotton) and negative if the CD decreases rst.
When measuring the CD spectrum, the Cotton eect generally
corresponds to the absorption maximum in the UVvis
spectrum, while the sign of the CD is determined by the
handedness of the supramolecular assemblies. If the enantiomers
are measured, mirror images should result.
Another technique is the exciton-type CD spectrum, as shown
in Figure 4A with dashed lines. This CD spectrum is commonly
referred to as a bisignate band, which is KramersKronig
consistent with the measured CB, as illustrated in Figure 4B.40 In
general, in the case of solutions, there is a direct correlation
between the regions of absorption and the CD. In the case of a
noncoupled chromophore, the shapes of the two spectra
observed for the enantiomers are similar, although the vibrational
ne structure could be dierent. If two or more strongly
absorbing chromophores are oriented chirally with respect to
each other, an exciton spectrum is observed and characterized by
the presence of two bands with opposite signs. The zero CD
between the valley and the peak is the crossover, which is usually
in the position of the absorption maximum in the UVvis
spectra. In supramolecular systems, the exciton CD is frequently
observed. While the exciton CD spectrum is generally symmetrical in solutions, it is quite common that the two bands of an
exciton spectrum do not have the same intensity in the
supramolecular system. This is because there are many
aggregated or cooperatively interacting chromophores, which
can lead to nondegenerate couplings of the same chromophore,
which alters the relative intensity of the two bands. In addition, in
some cases, the scattering eect may aect the shape of the CD
spectra.41

The transfer of chirality from a chiral center to aggregates or


assemblies is widely found in molecular self-assemblies. It is
generally accepted that chiral molecules easily form chiral
supramolecular systems. However, such a transfer, in fact,
depends on a number of factors including the distance of the
chiral center to the assembly site, the strength of the noncovalent
bonds, and the competition of the chiral and achiral interactions
to name just a few. The supramolecular chirality of the assemblies
can be determined by their CD spectral measurement and/or
morphological observation using AFM, SEM, and TEM.
Generally, in their monomeric or free state, the CD signal of
the compounds is silent in the chromophore portion if the
chromophore is located far from the chiral center. However,
through self-assembly, the entire assembly becomes chiral and
supramolecular chirality can be produced and thereby detected
by CD spectroscopy. During the self-assembly of the chiral
components, the chromophores are held together by various
noncovalent bonds and tend to adopt a spatial nonsymmetric or
chiral arrangement that lowers the energy of the system. In many
cases, such assemblies appear as chiral nanostructures, which can
be easily visualized using AFM or SEM observation. However, it
is not necessary that a chiral system that exhibits a CD signal
should always be composed of chiral nanostructures. The
chirality transfer will depend on the structure of the chiral
molecules. Here, we summarize the correlation between the
supramolecular chirality of typical chiral supramolecular systems
based on their component molecules.
4.1.1. Amphiphiles. Molecules that contain both hydrophilic and hydrophobic groups are called amphiphiles. A typical
amphiphile consists of a polar hydrophilic group, usually called
the head, which is joined to a nonpolar hydrophobic moiety,
referred to as the tail. Amphiphiles are some of the most
extensively investigated building blocks in self-assembly systems.
According to the number of polar head(s) and hydrophobic
tail(s) and the connection between them, amphiphiles are usually
classied as (1) single head/single or double tail amphiphiles, (2)
bolaamphiphiles, in which two hydrophilic heads are connected
by a hydrophobic skeleton group, (3) Gemini or dimeric
amphiphiles made up of two hydrocarbon tails and two ionic
groups linked by a spacer, or (4) dendritic amphiphiles, where
the headgroup or the hydrophobic chain is dendritic.
Amphiphilic molecules self-assemble in water or in an organic
phase to form various kinds of ordered structures including
micelles, vesicles, microemulsions, and liquid-crystalline mesomorphic phases. In addition, chiral self-assembly can hierarchically lead to a rich variety of more organized nanostructures such
as bers, ribbons, helices, superhelices, and tubes when the
amphiphiles are endowed with chiral elements. During this
process, supramolecular chirality is often, but not always,
expressed in the morphology of these aggregates at a length
scale of nanometers or micrometers.
4.1.1.1. Conventional Amphiphiles. Over the past three
decades, a number of synthetic chiral amphiphiles with diverse
molecular structures have been designed to form supramolecular
chiral nanostructures. In the design of amphiphiles, both the
position of the chiral center and the noncovalent bonding sites
should be considered. When chiral amphiphiles with long alkyl
chains self-assemble, a basic bilayer structure is generally formed
in the initial stage, which can further stack into multibilayer
structures. With the help of chiral sense in the amphiphiles, these
sheet-like bilayer membranes (single or multiple) often distort to

4. SUPRAMOLECULAR CHIRALITY IN
SELF-ASSEMBLED SYSTEMS CONTAINING CHIRAL
MOLECULAR COMPONENTS
During the chiral self-assembly, an important issue is how the
supramolecular chirality or the chiral nanostructures are
produced. Propagation of chiral information through specic
interactions and organization in materials or supramolecular
assemblies is generally called chirality transfer. The chirality
transfer from the molecules to the supramolecular system
represents an important origin of supramolecular chirality. This
chirality transfer in self-assemblies containing chiral molecular
components can be divided into two cases: (1) the chiral
information on a chiral center (generally an asymmetric carbon
atom or axial chirality) is imposed to the whole aggregate or
assemblies containing a chromophore, which usually can be
detected by CD spectroscopy, and (2) the chiral sense of one
component is transferred to achiral components to form a
complex system exhibiting supramolecular chirality. The bridges
for this chiral transfer are various noncovalent interactions, such
as hydrogen bonding, electrostatic interactions, metalion
coordination, donoracceptor interactions, hostguest interactions, and van der Waals interactions. In this section, we will
discuss the chirality transfer in systems composed of chiral
molecules and in those containing mixed chiral/achiral
molecules. Chiral systems composed exclusively from achiral
molecules will be the topic of the next section.
7310

DOI: 10.1021/cr500671p
Chem. Rev. 2015, 115, 73047397

Chemical Reviews

Review

to the superassemblies resulted from their dierences in the


hydrogen bonding between the gelator molecules. Both the 3b
and the 3c gelators self-assembled through strong intermolecular
H bonds and stacking, while 3a tended to form an
intramolecular H bond. Such dierences in the H bonding could
further lead to varying morphologies, in which the gels from 3b
and 3c formed nanotwists and nanotubes, respectively, exhibiting
a clear chiral feature in their morphologies, while 3a formed
conventional nanobers without a clear chiral feature.

form chiral nanostructures with a large curvature and/or a high


aspect ratio, leading to twists, helices, superhelices, tubes, and so
on. In 1984, Kunitake et al. reported the rst evidence of the
formation of nanoassemblies with a helical sense from a bilayer of
totally synthetic dialkyl amphiphiles.42,43
To date, a large number of amphiphiles have been reported to
form chiral brous aggregates based on distorted bilayers.44
These amphiphiles include diacetylenic phospholipids, glycolipids with open or cyclic sugars and various quantities and types
of unsaturation in the tail (double bonds, diacetylenic
moieties),45,46 amino-acid-based amphiphiles,47 as well as
hexamethylenediamine-based amphiphiles. More detailed reviews can be found elsewhere.13,4850 Herein, we show recent
examples of self-assemblies of the L-glutamic-acid-based
amphiphiles. Lius group and the Ihara group designed a series
of chiral amphiphiles based on L- or D-glutamide, and Liu et al.
investigated the chirality transfer from the molecular to
supramolecular level or nanostructures, as shown in Figure
5.30,5166 Self-assembled chiral nanostructures are generally

Figure 6. (A) Illustration of the self-assembly manner of dierent


isomeric molecules. Bilayer units were rst formed for 3a and 3c and
then stacked into multibilayers to form nanobers and nanotubes,
respectively. Only one bilayer unit is shown for clarity. 3b stacked into
square columns and then into a nanotwist. TEM images obtained from
various DMSO gels: (B) 3a, (C) 3b, and (D) 3c. Reprinted with
permission from ref 30. Copyright 2011 Royal Society of Chemistry.

4.1.1.2. Gemini Amphiphiles and Bolaamphiphiles. In


contrast to conventional amphiphiles with only one headgroup,
both Gemini amphiphiles and bolaamphiphiles have two
headgroups covalently connected by an alkyl spacer and, as
such, have attracted great interest in their manner of selfassembly and structure. During self-assembly, the cooperation of
these two headgroups together with the covalently linked spacer
plays an important role.
Oda et al. pioneered work on the helical self-assemblies
through gelation of cationic Gemini surfactants with chiral
counterions.6769 They found that the coassembly of cationic
Gemini amphiphiles with chiral tartrate counterions in chloroform leads to the formation of stable twist ribbons. L-Tartrate
produced exclusively right-handed helices, while the Denantiomer produced left-handed helices.
Bolaamphiphiles are molecules that contain two hydrophilic
groups covalent linked by a hydrophobic skeleton (e.g., one, two,
or three alkyl chains, a steroid, or a porphyrin).70 Bolaamphiphilic molecules have attracted much research interest due to the
fact that they can be found in archaebacteria, which are able to
survive in a volcanic environment, i.e., in hot sulfuric acid. It is
believed that the monolayer structure of a membrane together
with its helix formation can stien the cell membrane, enabling
the archaebacteria to survive in the hostile volcanic conditions.
The most important feature of assemblies formed by
bolaamphiphiles is their monolayer lipid membrane (MLM)
intermediate, which is very similar to a bilayer structure but the
two polar layers are covalently connected. The Shimizu research
group conducted a series of studies on the self-assembly of the
bolaamphiphiles that revealed their manner of assembly
systematically.71 During the self-assembly process, many of
these molecules formed into single- or multi-MLMs and then
self-assembled into nanotubes containing a variety of wall
thicknesses, lengths, and exterior/interior diameters.71 If the

Figure 5. Structure of glutamide amphiphiles (gelators) that formed


supramolecular chiral systems through gelation.

obtained by a supramolecular gelation procedure in which the


amphiphiles are dispersed into a solvent (water or an organic
solvent) by heating and/or sonication, and then the transparent
solution is allowed to cool to room or a designated temperature,
yielding supramolecular gels with an expression of supramolecular chirality and various chiral nanostructures.
For instance, three isomeric pyridine-containing (ortho, meta,
and para isomers) L-glutamic amphiphiles can easily form
organogels with DMSO (3ac).30 Although none of the gelator
molecules exhibited CD signals in solution, when they formed
the organogels, the assemblies based on the m-pyridinesubstituted (3b) and p-pyridine-substituted (3c) gelators
exhibited CD signals. However, the o-pyridine gels (3a) did
not. This dierence in the chiral transfer from the molecular scale
7311

DOI: 10.1021/cr500671p
Chem. Rev. 2015, 115, 73047397

Chemical Reviews

Review

and could be spun unto yarns using an automatic spinning


machine.75 The self-assembled supramolecular nanotube yarns
had a nominal tensile strength of 4560 MPa with Youngs
moduli as high as 6.89.9 GPa, which is comparable to many
covalently linked polymers. This result illustrates that a complex
supramolecular polymer with relative strong mechanical properties can be obtained through hierarchical self-assembly of a small
molecule. Here, use of the chirally terminated amino acids may
play an important role.
While the design of symmetric bolaamphiphiles is relative
simple, the self-assembly of unsymmetrical bolaamphiphiles
provides more diversity in the control of the inner and outer
surfaces of the resulting nanotube. Shimizu et al. designed
unsymmetrical L-glucosamide bolaamphiphiles that possessed
headgroups varying in size or properties, and these were used to
construct a lipid nanotube and an unsymmetrical lipid
membrane. Unsymmetrical MLM-based nanotubes possess
both an outer surface containing sugar hydroxyl groups and an
inner surface containing carboxylic acid groups. They were able
to control the inner diameters of the unsymmetrical MLM
nanotubes by varying the length of the spacer chain.76 Such
nanotubes oer two advantages when compared with those
prepared from the monopolar amphiphiles: (i) they possess
distinctive inner and outer surfaces, which can be used for
ecient encapsulation of materials, in particular, for selective
surface functionalization, and (ii) the diameter of these
nanotubes can be controlled by changing the molecular shape.
It is common practice to attach the chiral centers as the
headgroup when designing a chiral amphiphile. Thereby,
bolaamphiphiles with two headgroups of the same chirality are
generally designed. However, it would be interesting to observe
the chiral assembly process when the bolaamphiphiles contain
two headgroups with opposite chirality.
4.1.1.3. Peptide Amphiphiles. Among the many available
headgroups in an amphiphile, peptide amphiphiles (abbreviated
as PA) possessing peptides as the headgroup have attracted
increasing attention recently owing to their close relationship to
biological systems. Many types of peptides can be introduced
into the amphiphilic molecules ranging from simple amino acid
units to complicated peptide sequences with bioactive functions.
The group of Stupp77,78 developed a class of peptide
amphiphiles capable of self-assembly into cylindrical nanobers
with high aspect ratios. The chiral sense of the peptide oers an
opportunity to fabricate cylindrical intertwining nanobers that
form multiple helices or superhelices for biomedical applications.77,78
In the systematic study of peptide amphiphiles containing
valineglutamic acid dimers, Stupp et al. found that the dimeric
repeat unit promoted self-assembly into belt-like at assemblies.79 The lateral growth of these assemblies can be controlled
in the range from 100 nm to as little as 10 nm as the number of
dimeric repeat units increased from two to six. With the growth
of the peptide sequence, these at -sheet assemblies appeared to
twist (Figure 9). Experimental results demonstrated that the
peptide amphiphile sequences can profoundly aect the
subsequent supramolecular morphologies in water.
Of greater interest was the nding that the sequence of the
amino acids in the peptide amphiphiles could signicantly
inuence the one-dimensional (1D) chiral nanostructures. It was
found that four peptide amphiphile isomers, with identical
composition but varying sequences of four amino acids, had a
drastic eect on the resulting 1D nanostructures under identical
environmental conditions.80 The molecules with a peptide

headgroup is modied and the MLM is allowed to stabilize,


single-walled chiral nanotubes could possibly be fabricated from
the bolaamphiphiles. For example, Liu and co-workers designed
a simple bolaamphiphile containing two L-glutamic acids as the
hydrophilic head (15a, 15b) and found that this molecule can
self-assemble into a helical nanotube.27,72 The wall of the
nanotube consisted of a single molecular layer, and the
nanotubes were suciently long to form entanglements between
adjacent tubes. However, introduction of a rigid benzene
segment into the hydrophobic chain, in molecule 17, changed
the morphology of the resulting assemblies to nanobers or
nanoribbons.73 Another bolaamphiphile containing a methyl
ester of L-histidine (18) as the terminal group was also found to
form single-wall nanotubes at a low pH.74

Figure 7. Structures of bolaamphiphiles that were found to form chiral


nanotubes or twists.

Figure 8. AFM images of hydrogels of (A) 15c, (B) 18, and (C) 19.
Reprinted with permission from refs 72, 74, and 75. Copyright 2005 and
2013 Royal Society of Chemistry and Copyright 2013 John Wiley &
Sons.

Interestingly, when compound 18 was hydrolyzed to its


carboxylic acid, 19,75 it exhibited a hierarchical self-assembly into
multiwalled ultralong supramolecular nanotubes in slightly
alkaline aqueous solution (pH 89) as a result of hydrogen
bonding and electrostatic interaction. Of greater interest, these
nanotubes then formed bundles consisting of thousands of
ultralong supramolecular nanotubes packed in a parallel manner
7312

DOI: 10.1021/cr500671p
Chem. Rev. 2015, 115, 73047397

Chemical Reviews

Review

show strong CD signals in their absorption regions via selfassembly, which means the chirality can be transferred from the
periphery to the core and the whole assembly. A helical tube
fabricated from a dendritic glutamide lipid87 (21) exhibited a wall
with a thickness of 4 nm, which corresponds to a bilayer thickness
(Figure 11). While many other compounds generally form
multiwalled nanotubes, these dendron molecules appeared to
form a double-walled nanotube, because of the multiple H bonds
between the dendron heads that stabilize the bilayer structures.
More interestingly, the nanotube structure is very stable in a wide
pH range.
4.1.2. C3-Symmetric Molecules. Among the many supramolecular building motifs, C3-symmetric molecules have gained
special interest in forming organized nanoassemblies either
chirally or nonchirally.88,89 C3-symmetric building blocks, usually
consisting of a central aromatic or cyclohexane ring functionalized at the 1, 3, and 5 positions, have been frequently exploited.
In the design of the C3-symmetric molecules, other functional
groups can be attached via amide or urea bonds with tricarboxylic
acids or triamines. If chiral elements are introduced into these
molecules, chiral supramolecular assemblies can be easily
obtained. The C3-symmetric molecules often adopt a propellerlike conformation, because of the steric hindrance surrounding
the central aromatic core and the wedged substituent group
which causes rigidity through intramolecular interactions
(mainly hydrogen bonding). During stacking, the chiral elements
play an important role in determining the chiral sense of the
assemblies. The benzene-1,3,5-tricarboxamide (BTA) motif
comprising either three N-centered or three CO-centered
amides attached to a benzene core (Figure 12) has been widely
employed in the assembly of chiral supermolecules. The three
amide bonds in these molecules form H bonds, which contribute
to the one-dimensional growth of the monomers into a
columnar-type supramolecular polymer. The chiral information
at the substituents of the molecule can be transferred to the
central aromatic rings and induce preferential formation of a
single helicity in their supramolecular architectures. During this
process, chirality transfer is largely dependent on the size of the
C3 core and the connection between the wedged substituents
and the core.

Figure 9. (A) Structures of 20 (VE)2, (VE)4, and (VE)6. (BD)


Cryogenic transmission electron micrographs of (VE)2, (VE)4, and
(VE)6 nanostructures formed in 5 mM solutions of PA. Scale bars: 200
nm. Reprinted with permission from ref 79. Copyright 2013 American
Chemical Society.

sequence of alternating hydrophobic and hydrophilic amino


acids, such as VEVE and EVEV, are able to self-assemble into a
at nanostructure. EVEV was shown to form 3D twisted ribbons
or helical ribbons. In addition, occasional nanotubes can coexist
with the twisted and helical ribbons. By contrast, nonalternating
isomers such as VVEE and EEVV result only in the formation of
cylindrical nanobers that lack a chiral sense. This result is an
excellent demonstration that the peptide sequences can
determine the supramolecular chirality as well as the possible
chiral architectures (Figure 10).80
4.1.1.4. Dendritic Amphiphiles. Although dendritic amphiphiles have not been frequently reported in the past decade, they
exhibit some interesting assembly features with regard to the
expression of chirality.8186 For example, aromatic rings attached
to the core of a dendron with chiral glutamic acid substituents can

Figure 10. Structures of the isomeric peptide amphiphiles, and a schematic illustration of the nanostructures formed. Reprinted with permission from ref
80. Copyright 2014 American Chemical Society.
7313

DOI: 10.1021/cr500671p
Chem. Rev. 2015, 115, 73047397

Chemical Reviews

Review

Figure 11. Structure of dendron amphiphiles (left) and height AFM images of hydrogels of 21 (right): (A) pH 3, (B) pH 7, (C) pH 10, (D) pH 11. (E)
AFM image analysis of the helical pitch. (F) TEM image of a hydrogel obtained at pH 7. AFM and TEM images reprinted with permission from ref 87.
Copyright 2011 John Wiley & Sons.

NH bond compared to the PhCO bond in the monomer.


Therefore, N-centered BTAs may exhibit less cooperation in selfassembly relative to that of the CO-centered BTAs, which
results in a lower degree of amplication of chirality.101
To rationalize the chirality transfer mechanism in the selfassembly of C3 molecules, variations of the molecular structure,
including the distance of the chiral center to the central cores, the
number of chiral side chains, as well as the central core, have been
conducted. When replacement of a hydrogen atom by deuterium
was used as the source of chiral information, only a small energy
dierence between the diastereomerically related right- and lefthanded helical aggregates was observed (Figure 14).92 The
authors concluded that the value of the molar CD eect was
approximately three times lower than that of 34a, although the
sign and shape of the CD spectrum of the deuterated isomer 33a
was similar to that of 34a.
The structure in the central core has been found to have a
profound eect on the chirality transfer. Luis Sanchez et al.102,103
utilized oligo(phenyleneethynylene)-based tricarboxamides
(OPETAs) as the central core of the C3-symmetric discotics.
They compared the self-assembly of two series of C3-symmetric
discotics based on benzene-1,3,5-tricarboxamides (35) and
oligo(phenyleneethynylene)-based tricarboxamides (OPETAs)
(36), in which the peripheral groups were decorated with chiral
N-(2-aminoethyl)-3,4,5-trialkoxybenzamide units (35b and
36b) (Figure 15). Unexpectedly, the chiral BTAs here were
practically CD silent, which diered from previously reported
BTA derivatives. The authors speculated that the outer amide
functionalities of one molecule were too far apart to form
intermolecular, helical supramolecular polymers. However, the
CD spectrum of chiral OPETA 36b at a concentration of 1
105 M in MCH exhibited an intense bisignated Cotton eect,
which suggested that this compound had self-assembled into a
left-handed helical structure. These results conrmed the
combinatory inuence of the surface of the central aromatic
core and the branched nature of the peripheral side chains on the
chirality transfer.

Figure 12. (Top) Propeller-like shape of C3-symmetric molecules and


their assembly into a helical superstructure upon stacking in columns.
(Bottom) Structure of CO-centered and N-centered BTA. Reprinted
with permission from refs 101 and 222. Copyright 2010 and 2007 John
Wiley & Sons.

Meijer et al. presented pioneering work on the chiral selfassembly of BTA molecules.9093 The self-assembly of the parent
compound, CO-centered BTA equipped with chiral aliphatic
side chains, has been found to self-assemble into helical, onedimensional aggregates by means of strong, 3-fold H bonding,
which was conrmed by a strong Cotton eect centered around
220 nm. Later, several groups extended these studies by
synthesizing BTA derivatives with increased surfaces.
Examples include bipyridine,94,95 tetrathiafulvalene (TTF),96
NDI,97and porphyrins98,99 substituted with chiral side chains,
which were connected to the BTA core. Due to the large
conjugated surface, intermolecular self-assembly in solution takes
place via strong interactions and solvophobic eects.100
The self-assembly of N-BTAs shows self-assembly behavior
that is similar to that of their CO-centered counterparts, but
the amplication of chirality is less pronounced as the result of
the relatively weaker aggregation. A theoretical simulation reveals
that there is a higher energy penalty for rotation around the Ph
7314

DOI: 10.1021/cr500671p
Chem. Rev. 2015, 115, 73047397

Chemical Reviews

Review

Figure 13. Structures of CO-centered BTAs having intramolecular hydrogen bonds with increased surface.

Meijer et al. designed a C3-symmetric molecule that combined


the conjugated structures of OPEs with amide groups to
rebalance the stacking and H-bonding interactions.105 For
the chiral compound 40 at a concentration of 8 106 M in
chloroform, no CD eect was detected, suggesting that 40 was
molecularly dissolved at this concentration. However, in
methylcyclohexane (MCH), a bisignate Cotton eect was
observed, which indicated a preferred helicity in the columnar
aggregates formed by 40, induced by the optical activity.
Interestingly, when the authors changed the N-centered OPEbased discotics to CO-centered OPE-based discotics, the CD
spectra of the supramolecular polymers were opposite in sign
when keeping with the conguration of the identical stereogenic
center. This suggests that the helical preference of OPE-based
C3-symmetric molecules is governed by both the conguration of
the stereogenic center and the manner of the amide connectivity,
which aects the conformation of the amides with respect to the
-conjugated core. Furthermore, the authors evaluated the
impact of symmetrization of the discotic structure on the chirality
of the supramolecular polymer when C2-symmetric discotic trisamide 41 was synthesized. Despite 40 and 41 having an identical
conguration of the stereogenic centers, opposite chiral
information for the helical aggregates of 40 and 41 was observed.
Temperature-dependent CD measurements directly reected
weaker noncovalent interactions for the C2-symmetric 41 than

Two series of oligo(phenyleneethynylene) (OPE)-based C3


molecules have been synthesized to study the inuence of
structural features of the OPE discotics on chiroptical properties
(37, 38). Initially, the author compared the eect of the groups
linked between the aromatic core and the peripheral side chains
on the supramolecular chirality. The OPE-based trisamides with
a variable number of chiral side chains (compounds 37bd) selfassembled into helical aggregates. By contrast, the triangleshaped OPEs with ether and amide functional groups did not
exhibit eective chirality transfer (compounds 38), as demonstrated by the corresponding CD studies, even though there was
an absolute conguration of the stereogenic centers at all of the
peripheral chains. Therefore, the cooperation of the three highly
directional H bonds between the amide functional groups plays
an important role in the hierarchical self-assembly and
corresponding chirality transfer. Second, the CD spectra of the
OPE-based trisamides with a variable number of chiral side
chains demonstrated that only one stereogenic center was
needed to achieve a helical organization with a preferred
handedness. In addition, the ability to amplify the chirality
increased with an increase in the number of stereocenters at the
peripheral side chains. The study presented herein improved the
understanding of the structural rules that regulate the chiral
supramolecular organization of discrete molecules.104
7315

DOI: 10.1021/cr500671p
Chem. Rev. 2015, 115, 73047397

Chemical Reviews

Review

Figure 14. Structures of N,N,N-trialkylbenzene-1,3,5-tricarboxamides


(BTAs) (33, 34). Reprinted with permission from ref 92. Copyright
2010 Nature Publishing Group.

for 40. While the number of amide bonds was the same in both
systems, the strength of the hydrogen bonding varied. Combined
with the weaker stacking interactions in the C2-symmetric
discotics relative to the C3-symmetric discotics, the chirality
transfer may have occurred in a dierent manner.
The chirality of the chiral center can further be transferred to
determine the supramolecular chirality of the nanostructures.
For example, enantiomeric C3 compounds containing functional tetrathiafulvalene units (compound 31) selfassembled into helical aggregates showing a preferential helicity
twist over several length scales (Figure 17).96 The formation of
primary helices as twisted stacks was investigated by CD
measurements and further conrmed by the theoretical
calculations. molecular mechanics (MM) and molecular
dynamics (MD) simulations were used to evaluate the relative
stability of the P and M conformations of the stacks, and the
results indicated that the (S)-31 enantiomer provided the M
helix, which showed greater stability (2 kcal mol1) per molecule
than the P helix, which was in agreement with the optical activity
observed in the CD spectra. However, it was a surprise to obtain
mesoscopic-size chiral bers of this compound in higher
concentrations, which exhibited inverted helicity, i.e., P helices
for the S enantiomer and M helices for the R one. Although the
inversion of helicity between the primary twisted stack in
solution and the secondary helical aggregates for the solid bers,
which can be seen as superhelices from hierarchical assembly,
seems to be incomprehensible, it is a common phenomenon that
has been reported in many supramolecular systems. A more
detailed and exhaustive study needs to be conducted to explore
the link between molecular chirality, supramolecular chirality,
and the higher order of chiral expression or chiral nanostructures.
C3-symmetric benzene- or cyclohexane-centered chiral molecules often self-assemble into nanotube structures by column
stacking of the assembly unit. When chiral units are introduced,
the self-assembly process and formation of the nanotubes seems
to be easier. For example, a C3-symmetric L-glutamic acid ethyl

Figure 15. (A) Structures of C3-symmetric BTAs (35) and triangleshaped OPETAs (36). (B) CD spectra of compound 36b (2.5 106 M
in MCH) at room temperature and 90 C. Insets depict the cooling
curves of compound 36b from 363 to 288 K at intervals of 0.5 K min1.
Reprinted with permission from ref 102. Copyright 2013 John Wiley &
Sons.

ester gelator was found to form hexagonal tubes ranging from


nano- to micrometer scale depending on the solvents used.106 In
addition, through antisolvent gelation in a wide range of mixed
solvents, hexagonal nanotubes are formed instantly upon mixing
at room temperature.
4.1.3. -Conjugated Molecules. -Conjugated molecules
occupy a very important position in the research on supramolecular chirality of self-assembled systems. For this there are
several reasons. First, -conjugated molecules possess inherent
electronic properties. They are unique because of their potential
use in organic electronic devices, such as organic solar cells, eldeect transistors (FETs), light-emitting diodes (LEDs),
etc.107,108 If endowed with chirality, these molecules may
represent new structures with novel properties. Second,
stacking is one of the most important forms of noncovalent
bonding, which frequently determines if a system can perform
self-assembly and also determines the self-assembly pathway.109,110 Third, -conjugated molecules have strong absorption in the UVvis region, which allows their chiral assembly
processes to be easily characterized with CD spectra and other
morphological observations.
In order to realize chiral self-assemblies based on -conjugated
molecules, it is important to introduce chiral elements into the conjugated molecules. This can generally be done through
functional group substitutions at dierent positions on the
aromatic rings or in the substituted alkyl chains. In addition, the
7316

DOI: 10.1021/cr500671p
Chem. Rev. 2015, 115, 73047397

Chemical Reviews

Review

Figure 16. Structure of OPEs derivatives. Reprinted with permission


from refs 103 and 104. Copyright 2011 and 2012 American Chemical
Society.

chiral feature of the building unit itself, the substituent group, the
nature of the -conjugated backbone, solvent properties,
temperature, and even the stimuli factors such as light, heat,
sonication, magnetic elds, etc., can have a signicant inuence
on the chiral assemblies. Herein we demonstrate the typical chiral
assembly features for important -conjugated molecules such as
polycyclic aromatic hydrocarbons, thiophene and its derivatives,
oligo(p-phenylenevinylene) (OPV) and its derivatives, and
perylene bisimide (PBI) and its related compounds. Since the
self-assembly of a -conjugated molecule in a gel has been
reviewed,111,112 herein we focus on those materials with
supramolecular chirality.
4.1.3.1. Pyrene (a Polycyclic Aromatic Hydrocarbon). The
pyrene moiety is well known for its ability to form excimers at
specied concentrations in solution. In addition, the strong
stacking of the chromophore makes it easy to perform selfassembly, during which supramolecular chirality can also be
observed in the CD spectra if the chiral unit is introduced in the
vicinity of the chromophore. For example, compounds 42 and 43
contain a pyrene moiety covalently connected to the chiral unit
through the urethane moiety, forming organogels in isooctane
and n-dodecane, respectively. Although the solution of the
compounds did not show a CD signal, they exhibited CD signals
at gel states in the region of pyrene chromophores, indicating
that the chirality was transferred from the asymmetric carbon
atom to the whole assembly. A temperature-dependent CD
spectrum indicated that as the gel was progressively heated, the
CD signal decreased and completely disappeared as the gel
melted (Figure 18b), suggesting that chiral supramolecular
assembly was responsible for the CD signal.113
Two pyrene-conjugated glutamide derivatives were designed
to investigate the eect of the link spacer on the chirality transfer.
The pyrene moiety was linked to amphiphilic L-glutamide
directly (4a) or with three methylene spacers (4b). In both

Figure 17. Molecular structure of C3-symmetric compounds containing


tetrathiafulvalene units (31), and an SEM image of helical aggregates.
Reprinted with permission from ref 96. Copyright 2011 American
Chemical Society.

Figure 18. (a) CD spectra of isooctane gels of 42R (solid line) and 43R
(dotted line) at 293 K (1 mm path length). (b) Variable-temperature
CD spectra on an isooctane gel of 43R (31 mM). Reprinted with
permission from ref 113. Copyright 2010 American Chemical Society.

assemblies, the chirality in the L-glutamide moiety can be


transferred to the self-assembled nanostructures, but the
expression of the chirality at a supramolecular level appeared
to be dependent on the spacer and solvents. The 4b gels showed
the same P chirality both in polar and in nonpolar solvents;
however, the gel of 4a displayed an interesting chiral inversion
induced by solvent polarity, i.e., M chirality in nonpolar solvents
and P chirality in polar solvents (Figure 19). It was concluded
that the spacer between the amide groups and the pyrene ring
7317

DOI: 10.1021/cr500671p
Chem. Rev. 2015, 115, 73047397

Chemical Reviews

Review

Figure 19. Possible packing mode in the organogels of 4a (top) and 4b


(bottom). For the 4b organogel, the steric hindrance between the
pyrene moieties was reduced by the spacer. The units packed in the same
direction both in polar and in nonpolar solvents. For the 4a gel, similar
packing as in the 4b gel is observed in polar solvents. In nonpolar
solvents, due to the large steric hindrance and the stronger H bond
between the conjugated amide, opposite packing was adopted and the
chirality was inverted. Reprinted with permission from ref 60. Copyright
2013 Royal Society of Chemistry.

eectively regulated the hydrogen bonding and interactions


and then inuenced the assembly mode as well as the
corresponding supramolecular chirality.
4.1.3.2. Hexa-peri-hexabenzocoronene (HBC). Amphiphilic
hexa-peri-hexabenzocoronenes (HBC) developed by Aida et
al.114 are a unique class of amphiphilic aromatic molecules that
exhibit excellent self-assembly behavior. Although the achiral
HBC was found to self-assemble into chiral nanotubes, it
simultaneously formed two mirrored chiral assemblies of equal
quantity. Therefore, they canceled each other, and no macroscopic chirality could be detected. However, if some chiral
elements were introduced into the HBC molecules, a particular
supramolecular chirality can be obtained. For example, Aida et al.
found that a hexa-peri-hexabenzocoronene having two chiral
oxyalkylene side chains along with two lipophilic side chains
(denoted HBC 44) yielded graphitic nanotubes in MeTHF (2methyltetrahydrofuran).114 The CD spectra measurements
revealed that the solution of (S)-44 at 50 C was CD silent.
Cooling of the solution resulted in the appearance of positive CD
bands at 389, 400, and 423 nm, which gradually intensied with
time. This result suggested that the tubular assembled 44 most
likely contained a helical molecular arrangement of the -stacked
HBC units, whose helical sense was determined by the absolute
conguration of the chiral centers in the hydrophilic side chains.
Thus, the molecular chirality was successfully transferred into the
resulting supramolecular helical nanotubular assembly (Figure
20). By contrast, the HBC amphiphile 45 bearing branched
asymmetric centers in the paranic side chains produced few
nanotubular assemblies, which may be due to its branched
paranic side chains that prevent the formation of a bilayer tape
and further the chiral nanostructures. This indicated that not all
of the molecular chirality can be transferred to the supramolecular system.
4.1.3.3. Perylenebisimide (PBI). Perylenebisimide (PBI) is a
well-known dye which displays photostability and outstanding
optical and electronic properties as well as strong hydrophobic
interactions and stacking, which are of potential use in

Figure 20. Structures of HBC derivatives 44 and 45 and formation of


self-assembled graphitic nanotubes. (a) Schematic illustrations of the
structure of self-assembled graphitic nanotubes consisting of HBC
amphiphiles. (b) Formation of chiral graphitic nanotubes with onehanded helical arrays of -stacked HBC units through translation of
molecular chirality into supramolecular helical chirality. Reprinted with
permission from ref 114. Copyright 2005 National Academy of Sciences,
U.S.A.

electronic and photonic devices through the formation of selfassembled structures. There are many reports on the emergence
of supramolecular chirality based on PBIs as building units.115,116
For example, the aggregates of dipeptides and perylene
bisimide conjugates (glycine-tyrosine, GY, or glycine-aspartic
acid, GD) have been reported to show dierent chiral selfassembly depending on the nature of the peptide used. There is a
competition between H bonding among the peptides and the
aromatic stacking of PBI. Most interestingly, the peptide
sequence has a profound eect on the chirality transfer. In an
aqueous buer, PBI-[GY]2 formed chiral nanobers. In the
corresponding CD spectra of PBI-[GY]2 aggregates, a negative
Cotton eect at 447 nm and two positive Cotton eects were
observed at approximately 515 and 552 nm, indicating that the
chiral sense of the peptide was transferred to the PBI moiety. By
contrast, the PBI-[GD]2, which formed spherical aggregates
7318

DOI: 10.1021/cr500671p
Chem. Rev. 2015, 115, 73047397

Chemical Reviews

Review

rather than bers, was CD silent. This work established an


association of the chiral transfer from a molecularly chiral
component to supramolecular chirality as well as the morphology
of the assemblies. It further provides a strategy for using short
peptides and specically their sequence structure to manipulate
the PBI chiral nanostructure through chirality transfer.117

Figure 22. Chemical Structures of the S and R isomers of the compound


46 [R = CH(C6H13)2] and illustration of the self-assemblies structures
on the value of luminescence dissymmetry factor (glum). Reprinted with
permission from ref 119. Copyright 2014 American Chemical Society.

Figure 21. Circular dichroism spectra of PBI-[GY]2 and PBI-[GD]2


aggregates in buer solution (pH 10.8), concentration 1 103 M.
Reprinted with permission from ref 117. Copyright 2014 American
Chemical Society.

an even number of L-alanine groups in the side chains were


considerably less well dened and exhibited much smaller molar
ellipticities. Moreover, they observed a 2-fold oddeven eect,
i.e., on one hand, expressed by an alternating reversal of the
Cotton eect upon increasing the number of L-alanine units
within a series of molecules with the same spacer length. On the
other hand, the sign of the Cotton eect also alternated in reverse
with the increase in the spacer length for molecules with an
identical number of L-alanine units.
4.1.3.4. Phenylenes. Phenylenes are a typical class of rigid-rod
molecules that has been studied in the area of self-assembly as
well as optical properties. By introducing a chiral group in the coil
moiety of phenylenes, the chiral information can be transferred
to the phenylene moieties through the self-assembly process.121
For example, the bent rod-shaped rod molecule 47 was found to
self-assemble into the hollow tubules in dilute aqueous solutions.
Circular dichroism (CD) spectra of the aqueous solutions
showed a signicant Cotton eect above certain concentrations
(0.002 wt %) in the region of the aromatic chromophore,
indicating that the tubules adopted a one-handed helical
structure. Combining vapor pressure osmometry (VPO)
measurements and CPK models, Lee et al.122 proposed that
compound 47 self-assembles via a fully overlapped packing
arrangement into the hexamericmacrocycles, which, in turn,
stack on top of each other with mutual rotation in a single
direction to form helical tubules (Figure 23). When a pyridine
unit was introduced into the concave side of the apex of the bentshaped aromatic compound, 48, the pulsating motions of the
tubules were found to show a chiral inversion by virtue of the
pyridine forming water clusters through hydrogen bonding and
resulting in adjacent molecules that slide into a looser packing
arrangement.122
Another interesting example was taken from the helical selfassembly of oligo-p-phenylene-based organogelators 4952
(Figure 24), which has been found to be dependent on the

Many PBI helical assemblies are formed with the help of


intermolecular interactions because of their large planar
structure. However, Zhu et al.118 found that chiral carboxylicacid-functionalized PBI systems spontaneously self-assembled
into supramolecular helices via intermolecular hydrogen bonding
rather than stacking based on uorescence spectra
measurements. The uorescence of the PBI unit herein did not
change much upon formation of supramolecular assemblies,
which was in contrast to the arrangement, which led to
signicant uorescence quenching.
In addition to the chiral transfer from the chiral centers, the
axial chirality of binap was also found to be eciently transferred
to perylenebisimides (PBIs) with the help of molecular stacking.
The supramolecular chirality of the PBI assemblies was optically
probed by circular dichroism (CD), vibrational circular
dichroism (VCD), and circularly polarized luminescence
(CPL). The biPBI derivatives 46 formed one-dimensional
aggregates in methylcyclohexane (MCH) and spherical
aggregates in chloroform at a higher concentration.119 The
one-dimensional aggregates exhibited twice the value of the
luminescence dissymmetry factor (glum) when compared with the
spherical aggregates. The sum of excitonic couplings between the
individual chromophore units contributed to the high CPL
dissymmetry of the nanostructures.
Frauenrath and co-workers found an unprecedented 2-fold
oddeven eect during the investigation of oligopeptide
polymer-substituted perylene bisimides comprising a varying
number of L-alanines.120 Depending upon the number of Lalanine units, the observed CD activity was alternatively strong
and weak. The molecules bearing an odd number of L-alanines in
the side chains exhibited well-dened spectra with molar
ellipticities that increased directly with the number of L-alanine
residues. By contrast, the CD spectra of the compounds bearing
7319

DOI: 10.1021/cr500671p
Chem. Rev. 2015, 115, 73047397

Chemical Reviews

Review

Figure 23. (A) Structures of 47 and 48. (B) CD spectra of S-48 in aqueous solution at various concentrations. (C) Temperature-dependent CD spectra
of S-48 (0.01 wt %) in aqueous solution. (D) Schematic representation of reversible switching of the tubules between expanded and contracted states
with chirality inversion. Reprinted with permission from ref 122. Copyright 2012 American Association for the Advancement of Science.

Figure 24. Structures of 4952, and schematic illustration of the aggregation pathways. At low concentrations, low temperatures, and short times, the
helical organization of 50 is dominated by the atropisomerism of the central biphenyl unit and metastable P-type helices are formed. 51 and 52, and also
50 at higher concentrations, higher temperatures, and longer times, self-assemble into supramolecular structures of the opposite helicity (M-type).
Reprinted with permission from ref 123. Copyright 2013 John Wiley & Sons.

phenyl ring in the core.123 The OPPS with more than two phenyl
rings in the core self-assembled into left-handed helices, but
when it contained a biphenyl core the resulting molecule showed
exchangeable helicity depending on the reaction time, temperature, and concentration, exhibiting a competitive modulation of
supramolecular helicity controlled by a kinetic versus thermodynamic process. At low temperature, low concentration, and short
assembling times, compound 50 formed right-handed metastable
supramolecular helices under kinetic control. In this case, the
supramolecular chirality was determined by the axial chirality of
the biphenyl core. On the contrary, at higher temperature, higher
concentration, and longer reaction times, the external S

stereocenters determined the left-handed supramolecular helix


through thermodynamic control. By contrast, compounds 50
and 51, with three and four phenyl rings, respectively, selfassembled exclusively into left-handed aggregates, with no
inversion of the helicity. This indicated that the inuence of
the oligophenyl atropisomerism is much weaker than that of the
external stereocenters. This study provides not only a good
example of kinetically controlled modulation of supramolecular
chirality but also a better understanding of the chemical and
topological control in the generation of helical supramolecular
structures and the impact of the synergy between dierent chiral
elements.
7320

DOI: 10.1021/cr500671p
Chem. Rev. 2015, 115, 73047397

Chemical Reviews

Review

synthesized by Jiang et al.128 The self-assembly of this novel


porphyrinpentapeptide conjugate in THF/hexane and THF/
water was comparatively investigated to illustrate the eect of the
peptides second conformation on the helical arrangement of
porphyrin in the assemblies. The positive chirality of porphyrin
pentapeptide aggregates was observed from conjugates in THF/
hexane, suggesting the helical arrangement of the porphyrin
chromophore with a P helicity. The negative chirality was found
for the aggregates fabricated in THF/water, which was opposite
that observed in THF/hexane. This work conrmed that the
secondary conformation of the peripheral peptide tuned by
solvent polarity further inuenced the porphyrin chromophore
packing mode and supramolecular chirality in aggregates. This
result not only represented an example of organic nanostructures
self-assembled from a covalently linked porphyrinpentapeptide
conjugate but provided a strategy for controlling and tuning the
morphology and, in particular, the supramolecular chirality of
porphyrin nanostructures.
4.1.3.6. Oligo(p-phenylenevinylenes). Oligo(p-phenylenevinylenes) (OPVs) are another class of linear -conjugated
molecules which are widely used in the fabrication of chiral
supermolecules, due to the electronic properties of the system,
which is sensitive to intermolecular interactions, particularly the
way in which the chromophores are organized.
Ajayaghosh and co-workers reported many helical nanostructures composed of OPV assemblies that are formed by attaching
hydrocarbon chains with asymmetric carbons to the OPV
backbone, which have been reviewed recently.111 They mainly
focused on the emergence of supramolecular chirality during the
OPV organogelation, and a detailed description of the process
can be found in his review.
George and co-workers129 studied the self-assembly of OPVs
bearing a chiral side chain and obtained two dierent kinds of
assemblies which depended on the solvents employed and the
system temperature. Two aggregates were found, corresponding
to State A (2.5% THF) with sheet morphology and State B (10%
THF) with rolled nanotube structure. Remarkably, circular
dichroism (CD) studies performed on State B showed that these
assemblies were CD silent, while State A was marked by the
appearance of a bisignated CD signal with a positive Cotton
eect at 415 nm and a negative Cotton eect at 375 nm which is
characteristic of exciton-coupled OPV chromophores. The
chiroptical properties could be due to the dierent packing of
the chromophores. Notably, in State B, an annealing process was
found to improve the molecular ordering in the assemblies and
convert tubes to the nanosheets (similar to State A), which was
conrmed by a sudden appearance of the bisignated CD signals
during cooling, characteristic of excitonically coupled chromophores. Thus, the supramolecular chirality in the self-assembled
systems of the -conjugated molecules is strongly related to their
stacking even when they had chiral substituents.
Meijer, Schenning, and co-workers investigated the chiral
assembly of two OPVs through chiral peptide segments
composed of either a glycinyl-alanyl-glycinyl-alanyl-glycine
(GAGAG), silk-inspired -sheet, or a glycinyl-alanyl-asparagylprolyl-asparagyl-alanyl-alanyl-glycine (GANPNAAG), -turnforming oligopeptide sequence.130 Due to the dierent nature
of the two peptides, OPV-GAGAG dissolved molecularly in THF
and could only form a left-handed helix in water and MCH, while
OPV-GANPNAAG formed a left-handed helix in THF and
chloroform but a right-handed helix in water. In addition, the
stability of the formed chiral structures was remarkable. The
temperature-dependent CD spectra in water showed that the

4.1.3.5. Porphyrin. Porphyrin is one of the most extensively


investigated -conjugated compounds and exhibits excellent
assembly capability and biocomptability. Like other -conjugated
molecules, the introduction of chiral elements into the
macrocyclic ring induces chiral self-assembly and produces
supramolecular chirality. Porphyrin derivatives based on a
symmetrical amide-substituted discotic with chiral hydrocarbon
side chains were designed by Meijer et al.124 (Figure 25). At

Figure 25. (Top) Structure and temperature-dependent CD spectrum


of 53 in methylcyclohexane between 20 and 90 C with 10 C intervals.
(Bottom) A model in which aggregates, monomers, and monomeric and
dimeric porphyrin pyridine adducts are connected by equilibrium
constants. Reprinted with permission from ref 124. Copyright 2010
John Wiley & Sons.

room temperature, 53 was molecularly dissolved in chloroformand. It had a sharp Soret band at max = 422 nm but lacked a CD
signal. In MCH, porphyrin 53 exhibited a large blue shift to a
broadened band at max = 390 nm, which is a typical band
observed for cofacially arranged porphyrins or H aggregates.125
The CD measurements revealed an intense bisignate Cotton
eect with a crossover at 390 nm, indicating a helical
arrangement of the chromophores in the aggregate. The
porphyrin assemblies were disrupted by heating, accompanied
by a disappearance of the CD response. Upon cooling at a
concentration of 5.0 105 M, the CD eect reappeared at 69 C
(Figure 25).
A unique property of porphyrin compounds is their axial
coordination ability. Changes of the axial ligand will cause the
supramolecular chirality of the system to be regulated. For
example, the addition of the axial ligand pyridine to aggregates of
53 was found to alter the supramolecular chirality of the system.
Without pyridine, the porphyrin Soret band appears at 390 nm.
When 40 equiv of pyridine was added, a new red-shifted and split
band at 418 and 427 nm appeared. The exciton splitting energy
of 500 cm1 is indicative of a dimeric porphyrin pyridine
adduct.126 With the dissociation of porphyrin aggregates, the CD
spectrum in the exciton split band region became weak.
Ultimately, at a pyridine molar excess of 80 000, this split Soret
band gradually converted into a single, narrow, CD-silent band at
430 nm. This band was identical in shape and position to a
monomeric porphyrin pyridine adduct,127 which suggested that
at this rather high pyridine concentration, the porphyrin
aggregates have dissociated.
Another feature of the porphryin derivatives is their
modication in the central core by metal ions. A porphyrinato
zinc complex covalently linked with a peptide was designed and
7321

DOI: 10.1021/cr500671p
Chem. Rev. 2015, 115, 73047397

Chemical Reviews

Review

suprastructures with weak CD signals: this was positive for the


high molecular weight polymer and negative for the low
molecular weight polymer. On increasing the poor solvent in
the solution, the chirality of the high molecular weight polymer
showed no distinct changes in its CD spectrum, while for the low
molecular weight polymer, the single Cotton eect became a
bisignate Cotton eect with increased intensity. It was suggested
that the chiral centers can only oer small chiral perturbation to
the high molecular weight polymer systems but can impart helical
suprastructures to the low molecular weight polymer systems.
OPV has been incorporated into the main chain of chiral
poly(L-lactic acid)(PLLA), and the chirality transfer from PLLA
to OPV was investigated in the solid state.133 Under these
circumstances, the chirality information on PLLA was found to
be transferred to the self-assembled OPV chromophores. This
suggests that the molecular packing and supramolecular chirality
of OPV in the aggregates can be tuned by the PLLA. It was
observed that the mole percent incorporation of the OPV
chromophore can greatly aect the chirality transfer from PLLA,
and 3 mol % proved to be the incorporation ratio where the
strongest CD intensity was observed.
4.1.3.7. Oligo- and Poly(thiophenes). Thiophene and its
derivatives are another class of -conjugated molecules whose
chiral packing has been extensively studied. Chiral sexithiophenes can form helical aggregates in water, butanol, and the
solid state. In water and butanol, the chiral assemblies show
dierent melting transition temperatures. Furthermore, chiral
sexithiophenes have been found to induce the chiral packing of
achiral sexithiophenes. Interestingly, both thermodynamically
stable and kinetically favored mixed aggregates with opposite
supramolecular chirality were obtained. Later, Schenning and coworkers investigated the oddeven eect of the chiral center
position on the oligo(ethylene oxide) chains away from the
thiophene backbone and the number of thiophene rings on the
supramolecular chirality of the assemblies (Figure 27).134 It was
found that bisignate Cotton eects were observed for all these
molecules. As expected, the sign of the Cotton eect was reversed
for the aggregates of T6S and T6R due to the opposite
conguration of the stereocenter. Notably, the Cotton eect
showed positive signs for T5S and T7S but a negative sign for
T6S. In addition, when the chiral center was moved from the
to the position, there was a positive CD signal for T6S and
T6S and a negative CD signal for T6S and T6S. Therefore,
both the number of thiophene moieties and the chiral center on

helical aggregates of OPV-GAGAG were completely destroyed at


above 20 C, while CD could still be observed at 90 C for OPVGANPNAAG. The chiral assembly of OPV-peptide conjugates
depended greatly on solvent polarity, temperature, and, in
particular, the peptide nature. This was a good example of
regulation of supramolecular chirality of -conjugated systems
using peptide secondary structures.
Oxadiazole-containing OPVs have been chemically attached to
an -helical peptide, and the eect of the relative spacing and
orientation of the chromophores in the peptide on the chiral
assembly of the OPVs was explored (Figure 26).131 Oxa-6D

Figure 26. (Left) Structures of Oxa-OPV, Oxa-6D, Oxa-7O, and Oxa11D. (Right) CD spectra of Oxa-6D, Oxa-7O, Oxa-11D, and Oxa-6D
mono. A representative absorption spectrum for Oxa-11D is also shown.
Reprinted with permission from ref 131. Copyright 2008 American
Chemical Society.

mono (equipped with a single Oxa-OPV) showed no excitoncoupled CD signal. Oxa-6D and Oxa-7O, both of which have two
Oxa-OPVs with dierent orientations, generated very similar
negligible positive split Cotton eects. The weak Cotton eect
might be due to the high degree of overlap of the side
chromophores for Oxa-6D and the isolated Oxa-OPVs for Oxa7O. Oxa-11D equipped with two chromophores on the same side
with a lager spacing than Oxa-6D showed negative Cotton eects
with the greatest intensity among the three molecules. This
strong CD intensity suggested that the side chains of Oxa-11D
were in sucient proximity to one another for exciton coupling.
Chiral assemblies obtained from OPV polymers have also been
investigated.132 It was found that the chiral polymers (Rac and R)
with both high and low molecular weight could form helical

Figure 27. Structures of chiral oligothiophenes, and (a) CD spectra of T5S, T6S, and T7S in butanol (2.6 105 M) at 283 K. (b) CD spectra for all
chiral T6 derivatives (8 105 M) at 283 K. Reprinted with permission from ref 134. Copyright 2006 American Chemical Society.
7322

DOI: 10.1021/cr500671p
Chem. Rev. 2015, 115, 73047397

Chemical Reviews

Review

Figure 28. Inuence of the addition of methanol to a chloroform solution of the block copolymers 5457. (AD) CD spectra of compounds 5457 in
mixtures of methanol and chloroform. Reprinted with permission from ref 138. Copyright 2010 American Chemical Society.

have dierent eects on the self-assembly of chiral polythiophenes.


Thiophene block copolymers equipped with a chiral side chain
on one or both of the blocks were synthesized and investigated to
determine which chiral side chain played a key role in the chirality
transfer. Koeckelberghs et al.138 synthesized P3AT(S*)-bP3AOT, P3AT(R*)-b-P3AOT(S*), and P3AT(S*)-b-P3AOT(S*) (5457) composed of an alkyl- and an alkoxy-substituted
polythiophene block and investigated their aggregation behavior
using UVvis, CD, and emission spectroscopies. Through the
introduction of poor solvent, the chiral aggregation of the
P3AOT block occurred rst owing to its lower solubility in the
chosen solvent than that of P3AT. Upon further addition of the
poor solvent, the P3AT block also aggregated, thereby adopting
the same helical supramolecular organization as the P3AOT
block. The results obtained from CD spectroscopy suggested
that the P3AOT block transferred its helical supramolecular
structure to the P3AT block, because the chiroptical behavior of
P3AT(S*)-b-P3AOT signicantly contrasted with that of the
other three polymers. The achiral P3AOT block aggregated in an
achiral way regardless of the chirality of the P3AT block.
However, the substituent on the P3AT unit can complicate the
stacking, as expressed by the intensity of the CD spectra. These
studies revealed that for all block copolymers the initial block
aggregation addition of a nonsolvent has a major inuence on the
stacking and the chiroptical behavior of the other block.
4.1.3.8. Alkynylmetal. Yam et al.139 reported an example of
the control of the chiral supramolecular structures exerted by
variation of the counteranions in a single gelator molecule of a
luminescent chiral alkynylplatinum(II)terpyridyl. Through

the oligo(ethylene oxide) chains eected the chiral assembly in


an obvious oddeven manner.
Bauerle et al.135 explored how the chirality of the biomolecules
aected the chiral self-assembly of oligothiophenes. Initially, he
decorated the tetrathiophenes with carbohydrates, which
resulted in the chirality of the carbohydrate directing the chirality
of the assemblies of the carbohydratethiophene. The selfassembly of thiophenes into chiral superstructures can be tuned
by the choice of saccharidic building blocks with suitable
stereochemistry. The authors further synthesized thiophenes
containing a single chiral amino acid (proline) to study the chiral
assembly of the conjugates.135 It was demonstrated that proline
with two chiral centers could induce a dened helical packing of
the conjugates, whose helicity was controlled by the conguration of the amino acid moiety. Since proline has two chiral
centers, the authors also studied how the proline with opposite
chirality aects the self-assembly of prolinethiophene.136 The
thiophene functionalized with a proline unit of opposite chirality
showed a reversed Cotton eect in the region of the *
transition of thiophene, which indicated that the chirality was
transferred to the thiophene stacks and that the supramolecular
chirality was related to the stereochemistry of the proline residue.
The thiophene containing diastereomeric proline showed a silent
CD spectrum due to the lack of chirality in the formed
aggregates. The mixture of the two enantiomers showed a CD
spectrum in which the CD signal nearly vanished.
Besides thiophene oligomers, polythiophenes have also been
attracting much attention. Inganas et al.137 investigated the chiral
assembly of polythiophenes with synthetic peptides. In this work
it was observed that positively and negatively charged peptides
7323

DOI: 10.1021/cr500671p
Chem. Rev. 2015, 115, 73047397

Chemical Reviews

Review

metalmetal and interactions, molecule 58 can form


metallogels in DMSO and CD signals were obtained in the gels.
This result together with the lack of activity in the CD spectrum
of a solution of 58OTf in dichloromethane at the same
concentration indicated that the CD signals observed in the
metallogels originated from the helical chirality of the selfassembled chromophores, transferred from the chiral group,
rather than from the intrinsic chirality of the gelator molecules.
The chiral supramolecular structures of the metallogels are
inuenced by varying the counteranions. The metallogels with
various counteranions in DMSO showed dierent CD spectral
patterns with a negative Cotton eect in the region of the
MMLCT transition. The CD signal of 58OTf was much
stronger than that of other counteranions. The authors inferred
that 58OTf formed highly ordered helical supramolecular
structures. The dierent CD spectral patterns of 58 associated
with the dierent counteranions suggested that the variation of
the counteranions would give rise to dierent chiral supramolecular structures in the gel phase by varying the degree of
aggregation through PtPt and interactions.
Yam et al.140 also incorporated alkynylplatinum(II) terpyridine units into the single-turn backbone of a binaphthol
derivative, 59 (Figure 29). The complex experienced a transition

centers (heterochiral) can aect the supramolecular chirality of


the assemblies is a very interesting consideration. The important
question here is which chiral center will control the supramolecular chirality? Homochiral, heterochiral, and achiral
peptide auxiliaries appended with naphthalenediimides (NDIs)
were designed and synthesized, as shown in Figure 30. It was

Figure 29. Structure of alkynylplatinum(II) derivatives 58 and 59.


Reprinted with permission from refs 139 and140. Copyright 2009 John
Wiley & Sons and Copyright 2013 National Academy of Sciences,
U.S.A.

Figure 30. Structures of compound 60, homochiral (LL and DD),


heterochiral (LD and DL), and achiral (AA) peptide conjugates of NDI.
Proposed models: (a) Schematic representation of left-handed (LL and
LD) and right-handed (DD and DL) chiral supramolecular assemblies; (b)
schematic illustration of sergeants-and-soldiers eect in which the
achiral soldier AA follows the chiral sergeant, LL or DD. Reprinted with
permission from ref 141. Copyright 2012 John Wiley & Sons.

from random coils to single-turn helical strands in which the


conformational transition was controlled by the PtPt and
interactions of alkynylplatinum(II) terpyridine moieties based
on the solvents used and temperature. The bisignate Cotton
eect in the circular dichroism spectra was indicative of the
cooperative transformation from a random coil state to a
compact single-turn M or P helix. The metalmetal and
interactions of the alkynylplatinum(II) terpyridine moieties were
supposed to stabilize the metallofoldamers, as dened by density
functional theory calculations.140
4.1.4. Molecules with Multiple Chiral Centers. Thus far,
supramolecular chirality based on molecules with one or multiple
homochiral centers has been the central topic of this review.
However, how a molecule with two or more opposite chiral

found that in the case of the heterochiral peptide conjugates (LD


and DL) the chirality of the rst stereocenter (irrespective of the
stereochemistry of the second stereocenter) adjacent to the NDI
core determined the supramolecular helicity. Remarkably,
homochiral LL and DD peptide-modied NDIs self-assembled
into 1D hierarchical supramolecular polymers with opposite
helicity, while the heterochiral peptide conjugates LD and DL
formed microspheres.141
Yang et al.142 reported an interesting control of handedness of
the alanine dipeptide self-assemblies by the chirality of the
7324

DOI: 10.1021/cr500671p
Chem. Rev. 2015, 115, 73047397

Chemical Reviews

Review

alanine moiety adjacent to the hydrophobic headgroups.142 The


peptide amphiphiles termed (L,L)-61 and (L,D)-61 showed
negative CD signals, and (D,D)-61 and (D,L)-61 exhibited
opposite signals. Meanwhile, the handedness of the nanoribbons
of (L,L)-61 and (D,L)-61 were left-handed. The nanoribbons of
(L,D)-61 and (D,D)-61 were right-handed. The morphologies of
the assemblies were consistent with the CD spectral data, which
also indicated that the handedness of these organic selfassemblies was controlled by the chirality of the alanines at the
terminals.

often used for translation of chiral information from one


component to others in multicomponent assemblies.143149
For example, the pyridine-ended OPV 62 is a pregelator,
which can form supramolecular organogels via molecular
recognition of the chiral forms of tartaric acid (TA).147 The
molecule 62 with M-TA (meso-tartaric acid) did not form a gel,
suggesting that the spatial orientation of the COOH moiety in
the chiral TA played a key role in determining the gel formation
mode. The chiral information on the TA enantiomer led to the
formation of P- and M-helical bers in complexes containing DTA and L-TA, respectively, while both helices existed in the
complexes upon induction by M-TA. This result was consistent
with CD spectra measurements, in which the combination of 62
and enantiomeric TA revealed a bisignated Cotton eect. A
virtually mirror-image spectrum was observed for the complexes
with two enantiomers (L- and D-TA), indicating the transfer of
the chiral information on TA to the self-assembled chromophores in a helical sense. The complexes containing M-TA
showed no helical bias.
Chiral diamines were successfully employed as triggers to
transfer their chiral information to achiral tetracarboxymetallophthalocyanine 63 in DMSO/CHCl3 through hydrogen
bonding between carboxylic acid and amine.149 The sign and
amplitude of the supramolecular chirality was eected by the
structures of the amine molecules, volume ratio of the poor/good
cosolvents, type of poor solvents, molar ratio of chiral molecular
diamine to tcPcM, cavity metal of phthalocyanine, and addition
order of the amines.
Chirality transfer through hydrogen bonding between pyridine
and carboxylic acid was reported by the Liu group.150 As shown
in Figure 34, a simple supramolecular approach has been
proposed to achieve chirality transcription and resulted in
twisted nanostructures in a two-component system consisting of
L-glutamic-acid-based amphiphiles 64 and bipyridines 4Py.
Compound 64 can self-assemble into nanobers in water.
Upon coassembly with bipyridine, the nanostructures underwent
exciting changes to chiral twists due to strong hydrogen bonding
between the carboxylic acid and the pyridyl nitrogen atoms. The
molecular chirality of gelator molecules can be transferred to the
bipyridine aggregates by strong hydrogen bonding. Supramolecular chirality is expressed not only by the CD signals in
the corresponding absorption band of bipyridine but also by the
chiral twist structures.
De Feyter, Schenning, and Lazzaroni et al. reported the chiral
assembly of OPVs assisted by nucleobases and nucleosides.151153 Both achiral and chiral OPVs can form chiral
rosette structures. After addition of thymidine molecules, the
morphology of the OPVs transformed from rosettes to lamella
structures composed of dimers. The chirality of the lamella
depended on the chirality of the thymidine even for the chiral
OPVs. In addition, the OPVs can coassemble with thymine into
chiral patterns. It was found that the achiral guest molecule
thymine can induce the formation of diastereomers from an
enantiomeric OPV. The chirality of OPVs can then be tuned by
coadsorption with nucleobases or nucleosides.
4.2.1.2. Electrostatic Interaction. Electrostatic interactions
play an essential role in specic molecular recognition and
molecular assembly.154 An anionic chiral compound can induce a
cationic -conjugated polymer to form an interchain helically stacked assembly that is stabilized by both electrostatic and
interactions, which hierarchically self-organize into supermolecules with circularly polarized blue luminescence. A watersoluble poly(p-phenylene) derivative (PPP, 66) was synthesized

Figure 31. Molecular structures of compound 61; CD and UV spectra of


the hydrogels at a concentration of 30.0 g L1. Reprinted with
permission from ref 142. Copyright 2013 American Chemical Society.

4.2. Chirality Transfer in Systems Containing Chiral and


Achiral Molecules

Another important system of chirality transfer is that of a chiral


component to an achiral one and the extension over the whole
system. Here the induction of the chirality in the achiral
components is of utmost importance. In order to induce the
chirality of the achiral components, the interaction between the
chiral molecules and the achiral molecules plays a very important
role. Therefore, the design of molecules with matched bonding
sites and their cooperations are of utmost importance. All of the
noncovalent bondshydrogen bonds, electrostatic interactions,
hostguest interactions, as well as hydrophobic interactions
could be utilized to perform the chirality transfer between the
chiral and the achiral molecules. In some cases, chiral spaces or
environments can also endow achiral components with chirality.
An obvious merit of the chiral transfer in these system is that
instead of the tedious organic synthesis required to introduce
chiral units, a simple mixing of the functional achiral units with
the commercially available chiral molecules can produce
functional supramolecular chiral assemblies.
4.2.1. Chirality Transfer through Noncovalent Bonds.
4.2.1.1. H-Bond-Directed Chirality Transfer. Hydrogen bonding, which is directional and relatively strong, is the most
important interaction in self-assembly. Hydrogen bonding is
7325

DOI: 10.1021/cr500671p
Chem. Rev. 2015, 115, 73047397

Chemical Reviews

Review

Figure 32. Molecular structure of pyridine-end OPV 62, AFM topographic images showing (i) M helix for 62 + L-TA, (ii) P helix for 62 + D-TA, and (iii)
a mixture of M and P helices for 62 + M-TA. Reprinted with permission from ref 147. Copyright 2012 Royal Society of Chemistry.

Figure 33. Structures of tetracarboxymetallophthalocyanine 63 and chiral amines. Reprinted with permission from ref 149. Copyright 2011 John Wiley
& Sons.

of |10 2||101| in luminescence. Further, the polymer


assemblies gathered to form spherulites, which can be regarded
as semicrystalline nanospheres, and the spherulites exhibited
circularly polarized blue luminescence. This work provides a
simple way to fabricate chiral spherulites, which may nd
application in novel chiral nanomaterials for the next generation
of plastic optoelectronics.
Liu and co-workers reported that the enantiomer of
diaminocyclohexane induced a water-soluble porphyrin
(TPPS) to form helical nanorods in organic solvents.29 Mirrorimaged helical nanorods were observed in these systems,
indicating that the transferred chirality or induced chirality of
achiral -conjugated molecules can manifest not only via CD
spectra but also via direct helical nanostructures. Such chirality

by introducing tetraalkylammonium cations at the terminal sites


of the side chains, forming a complex with a water-soluble
diaxially chiral binaphthyl derivative (BNP) bearing two
sulfonate anions at terminal sites of the substituents, as shown
in Figure 36.155 A complex of 66 (PPP) and BNP exhibited a CD
band in the * transition region of the -conjugated backbone
(Figure 36a). The strong bisignate Cotton eects observed at
376 and 341 nm implied the presence of an exciton-coupling
phenomenon between the main chains. These results conrmed
that the assembly showed induced chirality of the polymer
moiety, which was caused by chirality transfer from the axially
chiral compound (BNP) to the achiral polymer (66). The
electrostatic interactions here are essential to the chirality
imposed on the polymers, resulting in large dissymmetry factors
7326

DOI: 10.1021/cr500671p
Chem. Rev. 2015, 115, 73047397

Chemical Reviews

Review

(CD), circularly polarized luminescence (CPL), optical


rotation), chiroptical switching processes, and nonlinear optical
(NLO) activity.
A supramolecular assembly of compound 68 (perylenebisimide (PBI) functionalized with dipicolylethylenediaminezinc
(DPAZn) binding sites), which can specically bind to
phosphates, showed an induced chirality when binding with
ATP (Figure 38).156 When DPAZn was mixed with 1 equiv of
ATP, a positive bisignate CD signal was observed, i.e., positive at
518 nm and then negative at 480 nm, in the PBI absorption
region with a zero crossing at 507 nm. This is characteristic of an
excitonically coupled right-handed helical organization of the
PBI chromophores, indicating the ecient chirality induction to
an achiral chromophoric assembly through the specic binding of
the phosphate guest molecules to the DPAZn sites.
Kleij described trinuclear Schi base host complexes in which
the conformation was rigidied by a central Zn ion.157 The
coordination of a series of suitable monotopic ligands to this
central Zn ion caused the eective chirality transfer to the host as
characterized by circular dichroism (CD) spectroscopy. The
chirality transfer provides the possibility for the development of
substrate-specic host systems that are useful for determination
of the absolute conguration of various types of organic
molecules.
Metalligand interactions were also attributed to express
chirality at the nanoscale via extending the -conjugated system
and enhancing the molecular interactions. Liu et al. reported that
the Cu2+ ions triggered the amphiphilic Schi base assemblies to
form twist nanobers.158 The square-planar coordination
between Cu2+ and the Schi base was attributed to the extension
of -conjugated system and further enhancement of the
intramolecular interactions, leading to the chirality being
expressed on the nanoscale, because the chiral interaction
accumulated in a conned space. Another example is a
terephthalic-acid-substituted amphiphilic L-glutamide (compound 6) gel in DMSO.57 A left-handed uniform helical twist
was obtained in the presence of a wide range of metal ions,
including Na+, Li+, Fe3+, Co2+, Ni2+, Cu2+, Zn2+, Cd2+, Mn2+,
Mg2+, Ca2+, Ag+, Eu3+, and Tb3+ (Figure 39). First, the ligand

Figure 34. Structures of compounds 64 and bipyridines (xPy).


Morphologies of coassembled 64/4Py (a, b, c), 64/4ePy (d), and 64/
2Py (e) at molar ratios of 1:2. Insets are photographs of the samples.
Reprinted with permission from ref 150. Copyright 2011 John Wiley &
Sons.

control over a large length scale from molecules to nanostructures could have implications in the design of asymmetric
nanocatalysts.
4.2.1.3. MetalLigand Coordination. The interactions
between metals and ligands are at the heart of a wide variety of
chemical, physical, and biological phenomena. Metalligand
interactions allow the design of materials with controlled
topology and with specic physical properties such as redox,
magnetic, or photochemical properties. Combined with a chiral
sense, metal-based materials can be designed that have unique
properties including chiroptical properties (circular dichroism

Figure 35. Structures of OPVs 65, thymidine, and thymine. Reprinted with permission from refs 151 and152. Copyright 2011 and 2013 American
Chemical Society. Reprinted with permission from ref 153. Copyright 2014 Royal Society of Chemistry.
7327

DOI: 10.1021/cr500671p
Chem. Rev. 2015, 115, 73047397

Chemical Reviews

Review

Figure 36. Poly(p-phenylene) derivative (PPP) 66 and binaphthyl derivatives (BNPs) with (R)- and (S)-congurations. A plausible model of the
electrostatic and interactions between two 66 repeating units and one (R)-BNP molecule. (a) UVvis absorption, CD, and gabs spectra, and (b) PL,
CPL, and glum spectra of PPP, BNP, and a mixture (PPPBNP) (1.0:2.0 mol/mol) in methanolwater (50:50 v/v). Reprinted with permission from ref
155. Copyright 2012 John Wiley & Sons.

molecules formed a at multibilayer structure through


stacking between the benzene rings and the H bonds between the
amide groups as well and the carboxylic acids. The interaction of
the metal ions with the carboxylic acid may change the at
multibilayer structure to a left-handed chiral twist due to the
chiral nature of the gelator molecules, providing an easy way to
tune the chiral twist by simply changing the metal ions.
4.2.1.4. HostGuest Interaction. The binding or encapsulation of a chiral guest in an achiral cavity has been proven to be an
eective method for chirality transfer. Rebek and co-workers
developed a hydrogen-bonded dimeric capsule composed of two
achiral monomers that produced a dissymmetric space in which a
small chiral guest can be encapsulated in a diastereoselective
fashion.159,160 This capsule was generally expressed as racemic,
because the enantiomers of the capsule are dynamically
interconvertible through dissociation and recombination. The
energy between the diastereomeric complexes would be dierent
as a result of the chiral interior recognizing the shape of the chiral
guest when a chiral guest is captured. The group of Haino
developed a calixarene-based capsule, which can encapsulate a
variety of guest molecules and heterodimeric hydrogen-bonded
pairs of carboxylic acids.161 When the chiral guest was

encapsulated, two diastereomeric isomers were formed,


suggesting that the P and M helicities of the capsule can be
biased by the chiral guest encapsulation.
A class of oligothiophene-based organogelator bearing two
crown ethers at both ends was found to gelatinize several organic
solvents in the presence of ammonium, forming one-dimensional
brous aggregates (Figure 40).162 The helical one-dimensional
assemblies were induced by the chirality of 1,2-bisammonium
guests through hostguest interactions. It was interesting to note
that the chirality of an oligothiophene-based organogel can be
created by thermal gelation, whereas it was silent in thixotropic
gelation.
4.2.1.5. Hydrophobic Interactions. Chirality transfer based
on hydrophobic interactions is rarely reported, which may be due
to the weakness of this interaction relative to other noncovalent
interactions. However, this is possible in gel systems where alkyl
chains of a chiral gelator and achiral guest molecules can entangle
each other to form chiral assemblies. We have presented chirality
transfer by taking advantage of this concept.51 The chirality can
be transferred to porphyrin chromophores through interchain
interaction between the alkyl chains of both the porphyrin and
the gelator during the coassembly. By contrast, when porphyrins
7328

DOI: 10.1021/cr500671p
Chem. Rev. 2015, 115, 73047397

Chemical Reviews

Review

Figure 38. Structure of molecule 68 (PDPA), and schematic of the


guest-induced regulation of supramolecular chirality in PDPA
assemblies. Reprinted with permission from ref 156. Copyright 2014
Royal Society of Chemistry.

Figure 37. (Top) Structures of TPPS (67) and chiral amines. (Bottom)
(A and B) Typical SEM images of the 67 (TPPS) nanostructures
assembled with the assistance of R-DAC (A) and S-DAC (B). (C) UV
vis (bottom) and CD (top) spectra of the dispersion of TPPS obtained
in the presence of R-DAC (black) and S-DAC (red). (DG) Typical
TEM and HRTEM images of the TPPS nanostructures assembled with
the assistance of R-DAC (D and E) and S-DAC (F and G). (Inset in
panels E and G) FFT of the corresponding HRTEM image. TPPS/DAC
ratio is 1/4. At the bottom is a schematic illustration of the formation of
mirror-imaged 67 nanorods with the assistance of DAC molecules.
Reprinted with permission from ref 29. Copyright 2013 Royal Society of
Chemistry.

Figure 39. Illustration of the self-assembly of 6 in DMSO. In the absence


of the metal ions, the bilayer structure was initially formed, many of
which further assembled into nanober structures. When metal ions
were present, they reacted with the headgroup and caused a twist of the
multibilayer structure. When Eu3+ and Tb3+ were added, red and green
emissive chiral twists were formed. For the sake of simplicity, only one
bilayer is shown. Reprinted with permission from ref 57. Copyright 2012
Royal Society of Chemistry.

without long alkyl chains were used in the gels, no CD signals


were observed in either the mixed solutions or the organogels.
This fact indicated that the entanglements or hydrophobic
interactions of long alkyl chains played an important role in
chirality transfer. In this case, the situation is more akin to achiral
molecules doped in the chiral liquid crystals. The chiral threedimensional microenvironment provided by the chiral gels is
believed to be responsible for the induction of the chirality of
TPPOC12H25 assemblies. Besides this approach, when an achiral
Schi base bearing long alkyl chains (70) was mixed with chiral
gels formed by 2L and 2D, the chiral information in the gelator
molecules was transferred to the Schi base chromophore and
supramolecular chirality was obtained.163 On the basis of the
dynamic covalent chemistry of the imine, the pH-responsive
property of the supramolecular chirality was explored and a pHdriven chiroptical switch was obtained upon treatment with acid
and base alternatively. Thus, a supramolecular chiroptical switch
was established based on supramolecular chirality transfer and
dynamic covalent chemistry.

An interesting example of the chirality transfer through


hydrophobic interactions was reported by Ghosh and coworkers. They investigated an H-bonding-mediated assembly
in bis-amide-functionalized chiral acceptor (NDI) and achiral
donor (DAN) molecules.164 Two types of homoaggregated
bers were obtained due to the mismatch in the distance between
the two amide groups. CD experiments revealed a helical
assembly for both the donor and the acceptor stacks, although a
chiral center was present only in the acceptor building block. The
authors suggested that the induction of helical bias was from the
acceptor stack to the donor stack via hydrophobic interaction
among the peripheral alkyl chains.
4.2.2. Chirality Transfer from Solvent to Assemblies.
The solvent, as the second supramolecular partner of each soft
self-assembly system, is crucial in determining the thermodynamic process of a self-assembly system. There are some
interesting cases that illustrate the chirality transfer from solvents
to supramolecular assemblies. The rst chiral solvent eect was
7329

DOI: 10.1021/cr500671p
Chem. Rev. 2015, 115, 73047397

Chemical Reviews

Review

Figure 40. Structure of oligothiophene derivative 69; CD spectra of the gel phase of 69 (straight line), 69(R,R)-diammonium (dashed line; triangle),
and 69(S,S)-diammonium (dotted line; circle). (a) ICD spectra of the 1:1 mixture. (b) CD titration study at 370 nm. Reprinted with permission from
ref 162. Copyright 2012 John Wiley & Sons.

OPV4UT in S-citronellol at room temperature showed a strong


bisignated Cotton eect, which is characteristic of the excitoncoupled helically ordered chromophore, similar to that reported
for homochiral OPV4UT analogues and the induced chirality
from citronellic acid guest molecules. This indicated the transfer
of chirality from the chiral solvent molecules to the racemic
stacks of achiral OPV molecules. The mirror image CD spectra
obtained for A-OPV4UT in the other enantiomeric chiral
solvent, R-citronellol, provided proof of chirality transfer. Similar
mirror-image, bisignated CD spectra and morphologies were
observed for A-OPV3UT in enantiomerically pure chiral
alcohols, showing that chiral induction in supramolecular stacks
through chiral solvents is possible.144
Zhang et al.172 synthesized an azo-containing -conjugated
polymer poly[(9,9-di-n-octyluorenyl-2,7-diyl)-alt-4,40azobenzene](F8AZO), 74, and found that the solvent chirality
of (S)- and (R)-limonenes was successfully transferred to mainchain polymers, which generated optically active 74 aggregates.
The intense circular dichroism (CD) signals corresponding to 74
in the visible region conrmed the chirality transfer from solvents
to polymer. More interestingly, the reversible chiroptical switch
was achieved upon alternating photoirradiation at 405 (trans
form) and 546 nm (cis form).
In general, the helical bias induced by a chiral solvent is not as
strong as the helical bias induced by a chiral monomer. For
example, chiral solvents (limonene) triggered assembly of a
racemic bisurea into a helical nanotube, which was characterized
by its circular dichroism signature. However, the helical bias was
only 33%, much lower than that induced by a chiral monomer.
However, this method oers a simple way to impose chirality on
supramolecular assemblies without introduction of a chiral
matrix or auxiliary.173
4.2.3. Chirality Transfer from Low Molecular Weight
Molecules to Macromolecules. While the synthesis of the
main-chain chiral polymers is an important topic, the regulation
or control of chirality of the polymer main chain through the

reported for the emergence of Cotton CD signals due to the


twisted form of CD-silent benzyl molecules dissolved in (2S,3S)butanediol.165 The second example was a helical preference
revealed by a study of the CD characteristics of poly(hexylisocyanate) in nonracemic chlorinated chiral solvents.166
A further study showed that helix formation can be induced in a
cosolvent containing chiral and achiral solvents.167,168
Recently, it was reported that the self-assembly of an achiral
perylenebisimide (PBI) organogelator (73) with two 3,4,5tridodecyloxybenzoylaminoethyl substituents at the imide
positions was chiroptically silent in achiral solvents. However,
in reality, it was found that this system formed both left- and
right-handed helices in equal amounts, which canceled any
chiroptical signal.169 When (R)- or (S)-limonene was used as a
chiral solvent, it was shown that a preferential population of a
certain handedness of the helical assemblies can be selected by
choosing a chiral solvent. Interestingly, the enantiomeric
selectivity depended on the assembly process. With dilute
solutions and sucient equilibration time (thermodynamic
conditions), the enantiomeric excess was close to 100%, whereas
for the assemblies fabricated by a controlled kinetic self-assembly
process (higher concentration), the enantiomeric excess was
only 20%. It was inferred that the fast gelation process at high
concentration was controlled by nonequilibrated nuclei in a
kinetic rather than a thermodynamic self-assembly process.
Under these conditions the chiral induction from the homochiral
solvent may not be adequate to eectively impose a single
handedness on helices.
The solvents not only induced the chirality transfer to
supermolecules assembled from low-mass molecules but can
impose the chiral information on solvents into the polymer or
oligopolymer assemblies.170,171 Achiral oligo(p-phenylenevinylene) (OPV) derivatives equipped with either ureidotriazine (AOPVUTs) or diaminotriazine (A-OPVTs) H-bonding arrays
were found to self-assemble into columnar stacks in apolar
solvents. When using enantiomerically pure R- and S-citronellol
as solvents, circular dichroism spectroscopy (CD) of A7330

DOI: 10.1021/cr500671p
Chem. Rev. 2015, 115, 73047397

Chemical Reviews

Review

chiral small molecules. Various acyclic and cyclic, primary,


secondary, and tertiary amines were introduced in the side chains
of the poly(acetylene)s. The same acidbase chemistry as above
holds for the interaction of these polymers with chiral carboxylic
acids. In addition, the resulting optical activity is strongly related
to the structure of the amine-functionalized polymer.
Other noncovalent interactions, such as metalligand
interactions, hostguest interactions, and electrostatic interactions, also contribute a great deal to inducing the formation of
one-handed helices from achiral polymers. For example, when
crown ethers were present in the pendants,184 a predominantly
one-handed helical conformation was formed in aqueous
solutions (HClO4) upon complexation with various chiral
compounds, such as amino acids, peptides, amino, sugars,
amines, and amino alcohols. Addition of NaCl or KCl decreased
the magnitude of the CD spectra, which suggested the
importance of the crown etherammonium complexation in
acidic water.
Interactions between chiral molecules and the functional
group in the polymer side chains were also observed for other
polymer backbones. Other achiral polymers, including polyisocyanates, poly(phenylisocyanide), poly(thiophene), and poly(guanidine), can also be endowed with dynamic chirality by
various noncovalent interactions.
Vandeleene et al. reported poly(phenyleneethynylene-altbithiophene) copolymers with chiral pendants and pendants
bearing carboxylic acid groups in solution and in lms.185 Here, it
was noticed that the addition of chiral primary amines resulted in
chiral aggregation of the polymers. When the chiral centers of the
pendants and amines were the same, cooperation between them
in helical stacks was observed by CD spectroscopy. The opposite
situation holds when they are opposed.
Inoue and co-workers synthesized a poly(m-ethynylpyridine)
polymer that comprised at least 72 pyridine moieties with a
molecular weight of ca. 4500.186 When a chiral saccharide was
enclosed in the inner sphere of the polymers, helical structures of
polymers were guided by uncharged hydrogen-bonding
interactions with saccharides. Circular dichroism studies revealed
the nature of the chirality induction and how the achiral host
senses dierently the chiral structure of a range of saccharides.

Figure 41. Structures of (A) dopant Schi base compound 70 and (B)
2L and 2D. (C) Schematic illustration of their coassembly. (a) A Schi
base based on a dynamic covalent bond. (b) In the coassembly of 70 and
2, 70 can be inserted into the alkyl chains of 2 molecules. On the basis of
hydrophobic interactions, a supramolecular assembly with twist
structure can be formed, and the supramolecular chirality can be
transferred from 2 to the Schi base moiety. The supramolecular
chirality showed on and o states through the alternate treatment of
acid and base. Reprinted with permission from ref 163. Copyright 2013
Royal Society of Chemistry.

interaction with a chiral unit in their side chain provides a new


arena for study (Figure 46).
The pioneering studies in this eld have been reviewed
thoroughly by Yashima et al.174 The rst helical polymer induced
by chiral amines through acidbase interactions was reported at
1995.175 Upon interacting with chiral amines in DMSO, a
preferred helical handedness of a cis-transoidal, stereoregular
poly((4-carboxyphenyl) acetylene) (75-f) is instantaneously
induced in the polymer, showing a characteristic ICD in the conjugated polymer backbone region, indicating that the chirality
of the amines was imposed into the main-chain polymers.175
Later, this helical sense induction concept through noncovalent
chirality transfer was applied to the synthesis of a variety of
chirality-responsive PPAs by introducing a specic functional
group as the pendant group.176183 Yashima and co-workers
systematically investigated the formation of one-handed helices
from achiral polymers by acidbase complexation of chiral
amines, amino alcohols, and amino acids with organic acid
functions in their side chains.175180 During this process, the
rather irregular twist of the adjacent double bonds around a
single bond is transformed into a helical conformation with a
predominant handedness. This was conrmed by bisignate
Cotton eects in the polymer backbone absorption, which
showed a mirror-image relationship between two enantiomers of

4.3. Dynamic Features and Regulation of Supramolecular


Chirality

Dynamic exchanges and rearrangements of building blocks in


assemblies present challenges in supramolecular chemistry. This
is also true for supramolecular chirality. In contrast to a system
under thermodynamic control, which often exhibits a single,
simple assembly route, supramolecular chirality based on
supramolecular chemistry also shows the complexity and
diversity of kinetic direction. Various noncovalent interactions
may result in nonequilibrium self-assembly, in which structural
diversity is achieved by forming several kinetic products based on
a single covalent building block. The multiple available
interaction sites and the exibility of the interaction modes
make the supramolecular chirality dependent on the kinetics of
self-assembly.
Stupp and co-workers187 demonstrated that the preparation
protocols of the peptide amphiphiles self-assembling in water can
result in the formation of dierent supramolecular morphologies,
either long laments containing -sheets or smaller aggregates
containing peptide segments in random coil conformations. The
peptide amphiphiles (PA) were found to exist as monomers in
the good solvent HFIP and formed assemblies upon addition of
7331

DOI: 10.1021/cr500671p
Chem. Rev. 2015, 115, 73047397

Chemical Reviews

Review

Figure 42. Schematic illustration of chiral induction by helical neighbors. Reprinted with permission from ref 164. Copyright 2011 Royal Society of
Chemistry.

molecular chirality that occurred in process I can be transferred


to a positive one, which is similar to that of process II.
In comparison with molecular chirality, supramolecular
chirality can be easily altered by external factors such as solvent,
temperature, sonication, photoirradiation, redox potential, and
chemical additives. This provides an opportunity to regulate the
supramolecular chirality in self-assembled systems.
4.3.1. Solvents. Solvent is the medium for self-assembly
processes and can strongly inuence self-assembly via the specic
interactions between solvent and solute. The basic features of
solvents such as polarity, viscosity, and solubility for the solute
and other compounds could also aect the supramolecular
chirality of a supramolecular system. The majority of reports in
the eld of supramolecular chirality focus on the inuence of the
molecular structure on assembly and largely ignore the role of the
solvent. However, understanding how solvent properties
inuence chiral structures can help provide a deep understanding
of how supramolecular chirality is produced.
For example, an L-glutamate-based amphiphilic gelator bearing
an azobenzene segment 5 formed organogels that showed an
excellent photoregulated gelsol transition.56 It was found that
totally opposite CD signals were observed in DMSO and toluene
gels. The DMSO gel exhibited a positive Cotton eect, while a
negative Cotton eect was observed in the toluene gel, as shown
in Figure 49. The opposite Cotton eect obtained from dierent
solvents implied that the supramolecular chirality was reversed as
a result of dierent molecular orientations at the molecular level.
According to the results of XRD and temperature-dependent
UVvis spectroscopy, two kinds of molecular stacking models

nonsolventwater. Two peptide amphiphile assemblies of the


same composition were prepared in two dierent ways, as shown
in Figure 47.187 Although the two systems have the same HFIP
and PA content, clear dierences can be observed between the
CD spectra of solutions 1 and 2 (Figure 47b). Solution 1
exhibited a random coil CD spectrum, whereas the spectrum of
solution 2 had -sheet character. The presence of -sheets in
solution 2 can be rationalized by the fact that for this solution the
stock solution of PA1 in HFIP is initially added to water. In
solution 2, although the amount of HEIP reached 20%, these sheets do not disassemble completely, because there was no
transition back to the random coil conformation. These results
demonstrate that -sheet assemblies have high kinetic stability
and, once formed, do not readily disassemble. It is evident that
insights into the characteristic dynamics of a supramolecular
system can provide an ecient way to select the optimum
assembly pathway necessary for function.
Zhang and Liu investigated the aggregation of an anionic
porphyrin (TPPS) (compound 67) on the a cationic polypeptide
(poly(lysine)) controlled by dynamic assembly.188 Through
simple adjustment of the mixing sequences of TPPS with
poly(lysine), opposite CD signals from TPPS J aggregation were
obtained. When PLL was dropped into the TPPS solution
(process I), a negative CD signal was observed. By contrast, a
positive Cotton eect was found when TPPS was added to the
PLL solution (process II). The time scan of the CD spectra
revealed that process II was controlled by thermodynamics, while
process I was controlled by kinetics. The negative supra7332

DOI: 10.1021/cr500671p
Chem. Rev. 2015, 115, 73047397

Chemical Reviews

Review

Figure 43. Molecular structure of PBI derivative 73. AFM images of a lm spin coated (2000 rpm) from a solution of PBI 1 in (S)-limonene (c = 1 104
M) onto HOPG. (a and b) Height images. (c) Phase image. Scale bars in a and c correspond to 450 nm; the z scale in a and b is 9 nm. The statistical graph
of M and P helices is derived from image a. (ab and cd) Cross-section analyses of the bers. Reprinted with permission from ref 169. Copyright 2013
John Wiley & Sons.

Figure 44. Schematic illustration of the preferential chiral solvation in OPVUT self-assembled stacks. Reprinted with permission from ref 144.
Copyright 2011 Royal Society of Chemistry.
7333

DOI: 10.1021/cr500671p
Chem. Rev. 2015, 115, 73047397

Chemical Reviews

Review

Figure 45. Schematic illustration of unpolarized-light-driven chiroptical switching in CHCl3/(1R or 1S)/IPA (0.3/1.5/1.2, v/v/v). In 74-trans, chiral
aggregation manifested as a bright yellow and turbid solution, while nonaggregated 74-cis formed a yellow transparent solution. Reprinted with
permission from ref 172. Copyright 2011 American Chemical Society.

Figure 46. Schematic illustration of helical polymers obtained from achiral polymers induced by achiral guests and structures of poly(phenylacetylene)s,
poly(thiophene)s, and poly(phosphazene)s. Reprinted with permission from ref 174. Copyright 2009 American Chemical Society.

been suggested that the interaction between the pyridylpyrazole


headgroup and the solvents may subtly change the stacking of the
molecules and thus their self-assembled nanostructures. Thus, by
choosing appropriate solvents, a transition in morphology from
nanobers to chiral twists to nanotubes and to microtubes can be
achieved.
Solvent-driven morphological transitions may dominate
supramolecular self-assembly in many cosolvent mixtures.189
For example, Liu et al. found that the addition of a small amount
of water to organic solvents, either water miscible or immiscible,

have been proposed: the azobenzene groups packed face to face


( stacking), and the amino acids packed by forming H bonds,
as shown in Figure 49.
In the case of a pyridylpyrazole-linked L-glutamide organogelator,58 diverse nanostructures over a wide scale range from
nanober to nanotube and microtubes were obtained based on
the polarity of the solvent. The nanober, nanotwist, nanotube,
and microtube structures of 9 were obtained in toluene,
chloroform, DMF, and DMSO, respectively. Such morphological
changes can also occur with xerogels in the solvent vapors. It has
7334

DOI: 10.1021/cr500671p
Chem. Rev. 2015, 115, 73047397

Chemical Reviews

Review

Figure 49. (A) CD spectra of a gel of 5 in various states: DMSO gel ()


with a concentration of 4.3 mM, and toluene gel () with a
concentration of 5.0 mM. (B) Schematic illustration of molecular
packing in DMSO and toluene. Reprinted with permission from ref 56.
Copyright 2011 American Chemical Society.
Figure 47. Assembly of PA1 is dependent on the preparation protocol.
(a) Two PA1 solutions (50 g mL1) in 20% HFIP were prepared via
two methods that dier by the order in which pure HFIP and the PA/
HFIP stock solution were added to water. Even though both solutions 1
and 2 contain the same PA concentration and HFIP content, clear
dierences can be observed in CD (b) and DLS (c). Time-dependent
CD (200 nm) acquired on solutions 1 and 2, shown in the inset of panel
b, demonstrates the large hysteresis involved. Reprinted with permission
from ref 187. Copyright 2014 American Chemical Society.

handed helix in nonpolar solvents to a right-handed helix in polar


solvents.

Figure 48. (A) CD spectra of a PLL/TPPS (67) mixture in dierent


mixing sequences: (a) PLL was added to TPPS (process I) and (b)
TPPS was added to PLL solution (process II). (B) Time-dependent CD
spectra of the PLL/TPPS mixture. Illustration of the formation of TPPS
aggregate on polymer. The green block represents TPPS, and its charges
were omitted. (a) Pending-type aggregate in which one site of TPPS
binds on the PLL, while the other unit is stacked on the rst in a head-totail manner as a J aggregate. When less PLL presented in the solution or
PLL was added into TPPS, such aggregates were predominantly formed
and the process is a dynamic one. (b) Wrapping-type aggregation in
which every TPPS unit is wrapping around the polymer chain, while
these units formed head-to-tail stacks as J aggregates. Reprinted with
permission from ref 188. Copyright 2009 American Chemical Society.

Figure 50. Chemical structure of the PTCDIHAG amphiphile


molecule 81. (a, c) CD spectra and (b, d) TEM images of the
PTCDIHAG molecules 81 in dierent volume ratios of CHCl3/nC8H18 or THF/H2O. Reprinted with permission from ref 190.
Copyright 2011 Royal Society of Chemistry.

4.3.2. Temperature. As mentioned above, hydrogen


bonding is the most important interaction in self-assembly: the
strength of this interaction decreases with increasing temperature. Therefore, chiral assemblies based on hydrogen bonding
are especially sensitive to the adjustment of temperature. The
most popular instances of this are reported in the supramolecular
gels based on H bonds, in which CD signals are silent in solution
or in the monomer state as the system temperature is increased.
Supramolecular chirality is induced by the gel formation process.
In other words, gelation-induced supramolecular chirality is quite
sensitive to the temperature-regulated solgel transformation.192
Meijer and co-worker investigated a series of supramolecular
polymers based on C3-symmetric molecules.9095 In these
systems, temperature-dependent supramolecular chirality is

can trigger the formation of chiral nanostructures of a cationic


amphiphile (7).61 In ethanol, nanobrous structures without any
chiral sense evolved into the helical nanostructures, and
furthermore, the helical pitch could be tuned by the amount of
water present. In nonpolar solvents, helical tube structures were
produced upon the addition of water.
The supramolecular interactions of the PBI derivatives were
successfully modulated by solvents,190,191 which not only
induced a CD signal inversion but also the macroscopic
properties could be modulated by the solvent, from a left7335

DOI: 10.1021/cr500671p
Chem. Rev. 2015, 115, 73047397

Chemical Reviews

Review

widely found. In general, at higher temperatures, the molecules


exist in the monomeric state, in which the CD is silent. Upon
cooling, the chirality appears gradually during the self-assembly.
Thus, temperature-dependent CD spectroscopy is often
performed to illustrate the mechanism of self-assembly.
With the exception of the thermally dependent change of
chirality in the supramolecular assemblies based on hydrogen
bonding, a thermally reversible method for the inversion of
chirality was developed by changing of the lattice symmetry. A
thermally reversible inversion of chirality was discovered in
helical supramolecular columns formed by C3-symmetric selfassembling dendrimers based on dendrons connected at their
apex via trisesters and trisamides of 1,3,5-benzenetricarboxylic
acid.86 The authors demonstrated a change in the lattice
symmetry as follows: negative chirality for 2D and 3D phases
with triangular symmetry (columnar hexagonal) and positive
chirality for 2D and 3D phases with rectangular symmetry
(columnar rectangular and orthorhombic) which was attributed
to a thermally-induced inversion. This is the rst example of the
elucidation of the mechanism of reversible inversion of helical
chirality in supramolecular dendrimers. The structural changes
reported can be used to design complex functions based on
helical supramolecular dendrimers with dierent degrees of
packing on their periphery.86
4.3.3. Redox Eect Chirality. When metal ions or moieties
with variable valence states are inserted into a molecular building
block, the redox chemistry of this moiety will cause the variation
of supramolecular chirality.193,194 Liu et al. reported redoxresponsive chiral organogels based on a Cu(II)quinolinol
derivative (10).54 With the gels of the Cu(II)quinolinol
derivative, a positive Cotton eect at 280 nm and a negative band
at 240 nm with a crossover at 260 nm and a single positive band at
351 nm and a negative band at 475 nm were present, which
suggested that the chirality was transferred onto the metallogel
assemblies. Upon reduction of Cu(II) to Cu(I) by ascorbic acid,
the signals at around 475 and 351 nm disappeared, whereas the
exciton band at around 260 nm was retained. This result showed
that the reduction was mainly localized on the central Cu(II)
ions, whereas the stacking of the aromatic rings, which were
brought proximal to the amide groups, was not destroyed. Owing
to the formation of this dierent structure, the chirality could not
propagate the LL band; thus, the bands at 351 and 475 nm
disappeared. Subsequently, as the system was oxidized by O2, the
negative Cotton eect at around 475 nm and the positive band at
351 nm reappeared. These results revealed that the supramolecular chirality could be tuned by redox chemistry and a
redox-driven chiroptical switch could be realized.54
4.3.4. Photoirradiation. Photoirradiation is a noninvasive,
easy, and fast external stimulus that is often utilized to adjust the
structures of supramolecular assemblies.56,195199 Photoinduced
isomerization of azobenzenes,200,201 dithienylethenes,202 and
spiropyrans is the most often-used strategy for developing
phototriggered chiroptical switches. An azobenzene-linked
phenyleneethynylene bearing chiral groups (82) showed an
intense positive signal at 464 nm with two negative signals at 407
and 322 nm with a zero crossing at 421 nm corresponding to the
* transition of the PE moiety. Surprisingly, the CD spectrum
after UV irradiation at 323 K, followed by cooling the solution to
a lower temperature, showed a reversal of the CD spectrum;
however, there was a lower intensity in the CD bands. In this
case, photoirradiation resulted in a reversal of the CD signal.203
SEM analysis of (S)-83 before photoirradiation showed
entangled right-handed (P) helical ropes of diameters ranging

Figure 51. (Top) Cu(10)2 molecules self-assembled into gels through


coordination, hydrogen bonding, and hydrophobic interactions, as well
as through stacking. On reduction, this stacking was impeded,
and accordingly, the gels changed into a sol. (Bottom) CD spectra of 10
(a) and Cu(10)2 gels (b), the sol after reduction (c), and the revived gel
after redox (d). (Inset) Enlargement of the spectra in the range 400
500 nm. Reprinted with permission from ref 54. Copyright 2013 John
Wiley & Sons.

from 50 nm to 1 mm with lengths of several micrometers (Figure


52a). After irradiation at 323 K followed by cooling, the helicity
of the bers was found to be left handed (M), as shown in Figure
52b. AFM analysis of (S)-83 revealed reversal of their native
helicity to the induced opposite screw sense after irradiation.
In the absence of heating, the helicity did not change even after
irradiation for several hours. Even if there could be slow
isomerization of the molecules within the helical bers, the
helical twist did not change signicantly. This work demonstrated that the handedness of a photoresponsive supramolecular
object can be tuned with the cooperation of light and heat,
without changing the inherent molecular chirality of the
individual building blocks.
Cone-shaped alkoxyazobenzenes dimers functionalized with
amide groups were synthesized, and the presence of amide
hydrogen-bonding sites in one side of the folded molecules
prevented the antiparallel stacking favored by asymmetric
structures, facilitating the formation of toroidal aggregates, in
which the chirality was transferred to supramolecular chirality.
The resulting toroidal structures have large surfaces on their
top and bottom and can hierarchically organize into tubular
nanostructures. Irradiation of 84 with UV light quantitatively
converted aggregative trans-azobenzene moieties to nonaggregative cis isomers (Figure 53a),204 collapsing nearly all aggregates
to monomers (DLS study). However, a dierent situation was
achieved when 84 was irradiated with visible light at 470 nm: the
reversible transcis isomerization of the azobenzene moieties
could be promoted because both isomers have absorptions at 470
nm, while the fraction of the aggregative trans isomer is kept in
large excess because the cis isomer has a greater molar extinction
coecient at this wavelength (Figure 53b). In this situation, a
rapid generation and evanescence of polar cis-azobenzene
moieties occurred within the toroidal aggregates. This is capable
of promoting hierarchical growth of the aggregates by increasing
nanostructure surface polarity in nonpolar environments.
4.3.5. Chemical Additives. The addition of guest molecules
and metal ions was found to be eective in adjusting the chiral
superstructures of assemblies through guesthost interaction
7336

DOI: 10.1021/cr500671p
Chem. Rev. 2015, 115, 73047397

Chemical Reviews

Review

Figure 52. Photoisomerization of the azobenzene-linked phenyleneethynylene derivatives 82 and 83. A mixture of E,E, E,Z, and Z,Z isomers is possible.
SEM images of (S)-83 (a) before and (b) after photoisomerization; AFM images for (S)-83 (c) before and (d) after photoisomerization. Reprinted with
permission from ref 203. Copyright 2012 John Wiley & Sons.

and metalligand cooperation.57,158,205208 For example, an


achiral-guest-triggered chiral inversion in a novel supramolecular
assembly fabricated by pillar[5]arenes has been reported.209 The
planar chirality of pillar[5]arenes is caused by the substitution
position of the alkoxy groups, which have two equivalent stable
conformations (pS and pR), as shown in Figure 54B(a). In order
to isolate the two enantiomers, the rotation of these units should
be inhibited because the interconversion between pS and pR
occurs by rotation of these units in solution. General approaches
to hinder this rotation are to modify both rims with bulky
substituents or form a rotaxane consisting of a pillar[5]arene
wheel and a guest axle.

Yamagishi et al.209 reported a highly stable 1:1 hostguest


complex formed by pseudo[1]catenane 1,1,4-dicyanobutane
(G1) as a guest and pillar[5]arenes as a host. Inclusion of G1 in
the cavity of pillar[5]arenes causes dethreading of the alkyl chain
moiety from the cavity of pillar[5]arenes, which caused the
achiral guest G1 to induce the planar chiral inversion from in-pS1 to out-pR-1. The chiral inversion was characterized by the
observed CD spectra, which changed dramatically from positive
to negative with increasing G1 concentration. An ammonium
cation (G2) was another guest that could be added to the second
fraction (in-pR-1), and a decrease in the CD intensities was
observed. The authors concluded that the absence of chiral
inversion was due to the weaker association constants between
7337

DOI: 10.1021/cr500671p
Chem. Rev. 2015, 115, 73047397

Chemical Reviews

Review

Figure 53. (A) Schematic illustration of the self-assembly of 84. (B) Photoinduced (a and b) UVvis and (c and d) CD spectral change of 84 (c = 3.0
104 M) in MCH at 20 C. (a and c) Changes upon irradiation of a trans-rich solution with 365 nm UV light. (b and d) Changes upon irradiation of the
cis-rich solution with 470 nm visible light. (e) Plots of the fraction of cis-azobenzene moieties (red marks, left axis), and maximum values (blue marks,
right axis) versus irradiation time of UV (left side) and visible (right side) light. Aggregation states are shown with graduated background colors
representing monomeric (water blue) and aggregated (orange) states. Reprinted with permission from ref 204. Copyright 2012 American Chemical
Society.
7338

DOI: 10.1021/cr500671p
Chem. Rev. 2015, 115, 73047397

Chemical Reviews

Review

Figure 54. (A) Molecular structure of pillar[5]arene, competitive guests (G1 and G2), and a competitive host ([24,8]). (B) Representations of (a) the
planar chiral inversions triggered by achiral guest G1 and (b) alternating addition of achiral guest G2 and host [24,8]. Reprinted with permission from ref
209. Copyright 2013 John Wiley & Sons.

G2 and pillar[5]arenes than those between G1 and pillar[5]arenes. More interestingly, G2 could be removed from the cavity
of pillar[5]arenes with the assistance of the crown ether
[24]crown-8 ([24,8]), resulting in CD intensities that nearly
recovered to their initial state. This result displayed the host
guest complexation as a valid driving force for the chiral
inversion, and this guest-triggered chiral inversion system will be
useful for chiral switching or sensing systems.
Naphthalene diimide amphiphiles functionalized with the
dipicolylethylenediamineZn motif were synthesized in order to
promote a guest-induced self-assembly and chiral induction
through specic binding interactions.210 Titration of NDPA
amphiphiles with increasing molar ratios of ADP resulted in the
gradual evolution of strong Cotton eects, indicating that ADP
binding induced a preferred helical handedness to the resulting
assemblies of achiral NDIs. The binding of ATP induced
opposite handedness to NDI assemblies, as evident from the
positive bisignated CD signal, with positive and negative maxima
at 390 and 359 nm, respectively. The mirror-image Cotton
eects of NDPA-Bola assemblies obtained with ADP and ATP
indicated the induction of chirality with opposite handedness.
Interestingly, addition of 0.5 equiv of ATP to NDPA-Amph/
ADP assemblies resulted in positive bisignated CD signals which
exactly match with those of NDPA-Amph/ATP stacks alone.
This clearly suggested the competitive replacement of ADP by
ATP from the assemblies as expected and an instantaneous
reversal of its helical handedness. The authors also revealed a
dynamic helix reversal procedure through an intrastack
mechanism.
4.3.6. Sonication. Ultrasound is often used as a source of
energy to cleave and homogenize H-bonding, stacking, and
hydrophobic interactions of molecular building blocks and to
reshape the packing mode and the morphology.211216
The self-assembly of bichromophoric perylene bisimide into
chiral nanostructures, and the supramolecular helicity of the
nanostructures could be controlled by varying the method of

Figure 55. (a) CD spectra and (b) schematic of the dynamic helical
reversal of NDPA-Amph/ADP assemblies upon competitive guest
binding experiments with ATP (c = 7 105 M, 70% aq HEPES buer in
THF). Reprinted with permission from ref 210. Copyright 2012 Royal
Society of Chemistry.

preparation.217 The aggregates prepared by the heatingcooling


method possessed ordered molecular packing and enhanced
optical chirality. In contrast, ultrasonication resulted in molecular
aggregates with less ordered packing and opposite supramolecular chirality to the sample prepared via a heatingcooling
method. This heatingcooling method caused the nanobers to
have extended length and a prominent helical twist. The S isomer
gave left-handed M helices, and R isomers provided right-handed
chiral sense. In contrast, the assemblies prepared by the
ultrasonication method exhibited thinner bers (1015 nm)
with the opposite twist in helices for the corresponding isomers,
left- (M) and right-handed (P) twists for the R and S isomers,
respectively. The procedure which uses a heatingcooling cycle
is thermodynamically driven and results in the formation of more
stable nanostructures. In contrast, ultrasonication is a fast process
leading to a kinetically stable product in a shorter time frame. The
tunable chiroptical properties in these supramolecular systems
make them potential candidates for applications in the eld of
optical and electronic device fabrication based on organic
nanostructures.
7339

DOI: 10.1021/cr500671p
Chem. Rev. 2015, 115, 73047397

Chemical Reviews

Review

Figure 56. Schematic illustration of the dierence in molecular packing


leading to reversal of supramolecular chirality of the aggregates formed
by the two methods. Reprinted with permission from ref 217. Copyright
2011 American Chemical Society.

4.3.7. pH Value. The helicity inversion can often be found to


be triggered by the change of pH value through the change of
conformation of peptide and the interaction between building
blocks that are based on amino acids.218220
Lednev et al. reported that a small pH change initiated
spontaneous transformation of insulin brils from one
polymorph to another.220 These authors found that the sign of
the VCD band pattern from lament chirality can be controlled
by adjusting the pH of the incubating solution, above pH 2 for
normal left-hand helical laments and below pH 2 for
reversed right-hand helical laments. Later, they extended
this to other proteins and peptide fragments and again found that
pH variation triggered lament chirality change.

Figure 58. Illustration of chiral amplication and the model of two


principles: sergeants-and-soldiers rule and majority rule.

controlling the supramolecular chirality of the system, and in the


case of majority rule, how small can a chiral bias be while still
determining the chirality of an entire system.
In the rst case, a large amount of achiral units (the soldiers)
obey the rule of a small number of chiral molecules (the
sergeants). Majority rules refers to a slight initial excess of a single
enantiomer leading to a strong bias toward the same helical sense
in the whole aggregate.
4.4.1. Analogue-Induced Chiral Amplication. Meijer et
al.94 rst attempted to apply these two rules to explain the chiral
amplication in supramolecular systems in solutions. When a
small amount of a chiral disc-shaped molecule was added to a
solution of an achiral analogue in hexane, the CD eect of the
bipyridine transition showed a nonlinear response to the amount
of chiral sergeant.94,96,127,222 Fitting the data to a theoretical
model showed that on introducing on average one molecule of
chiral 32a per 80 molecules of achiral 32b, the chiral component
(the sergeant) dictated the helical sense of the total stack (of
soldiers). The mixtures of the enantiomers 32a and 32c showed a
nonlinear response of the CD eect on the enantiomeric excess,
indicating that chiral amplication in these systems corresponds
to the majority-rules eect.
Many of the supramolecular systems were found to obey these
rules. Coronenebisimides (CBIs), as potential candidates for
novel liquid-crystalline materials and active n-type semiconductor molecules in organic electronics, were assembled in
nonpolar methylcyclohexane.223 Derivatives of CBIs bearing
chiral and achiral 3,4,5-trialkoxyphenyl groups at the imide
position (85 and 86) self-assembled mainly through -stacking
and van der Waals interactions in methylcyclohexane, resulting in
long 1D brillar stacks. Dierent amounts of 85 were
coassembled with 86 (c = 2.5 105 M), and their chiroptical
properties were probed. Even with a small amount (3%) of the
chiral derivative (85) as the sergeant the CD spectrum of the
coassembly showed a bisignated Cotton eect (Figure 60a). The
anisotropy factor or g value / monitored at = 320 nm
showed nonlinear behavior (Figure 60b), which reached the
corresponding value of the pure chiral assembly at around 50% of
the sergeant. This result suggested a chiral amplication based on
sergeants-and-soldiers rule in this system.
The eect of chemical structure on the amplication of
chirality was studied by systematic variation of the chemical
structure of benzene-1,3,5-tricarboxamide derivatives 8792
(Figure 61).224 Since each BTA comprises three side groups,

Figure 57. AFM images of prion brils grown in pH 2.0 (a) and 3.9 (b).
Reprinted with permission from ref 220. Copyright 2014 American
Chemical Society.

4.4. Chiral Amplication in Supramolecular Systems

Amplication of chirality is a well-known phenomenon in


classical covalent polymers, the pioneering studies of which were
performed by Green and co-workers using the poly(alkylisocyanates) system.221 They dened two eects that
inuence the amplication of chirality and named them the
sergeants-and-soldiers principle and the majority-rules
eects. In recent decades, interest in the amplication of chirality
has broadened to supramolecular polymer systems based on
noncovalent interactions. In the supramolecular assemblies,
chiral amplication has been described as a phenomenon where
local chirality of a small fraction of chiral bias decides the chiral
sense of the entire assembly and is in general followed by
manifestation in the CD signals (Figure 58). The two principles,
sergeants-and-soldiers and majority rules, often describe the
strong amplication of a small chiral imbalance at the molecular
level to a supramolecular chirality. The basic concepts of
amplication of chirality in the self-assembled systems are
illustrated in Figure 58.
The key challenge in the case of the sergeant-and-soldiers rule
is how small can the amount of sergeant molecules be while still
7340

DOI: 10.1021/cr500671p
Chem. Rev. 2015, 115, 73047397

Chemical Reviews

Review

Figure 59. (a) Hydrophobic disc-shaped compounds with bipyridine units (32ac). (b) Amplication of chirality observed upon mixing solutions of
32a and 32b in hexane results in a nonlinear relationship between the CD eect and the amount of chiral 32a added to achiral 32b. (c) Net helicity as a
function of enantiomeric excess measured by CD spectroscopy of mixtures of 32a and 32c in n-octane. The line indicates the theoretical result that gives
the closest agreement with the experiment. Reprinted with permission from ref 96. Copyright 2011 American Chemical Society.

Figure 60. Coronene bisimide molecules 85 and 86 and coassembly of 85 and 86 and resultant chiral amplication. All experiments were done in MCH
(c = 2.5 105 M). (a) CD spectra of the coassembly at dierent percentages of 85 in a 1 cm cuvette at 20 C. The arrow indicates the spectral change
with an increase in the percentage of the sergeant. (b) Anisotropy value or g value monitored at = 320 nm as a function of the percentage of 85. The
dashed line that connects the fraction of 85 indicates the linear variation of the g value in the absence of any chiral amplication. Reprinted with
permission from ref 223. Copyright 2013 John Wiley & Sons.

soldiers experiments, it was found that 90 showed a stronger


amplication of chirality in majority-rules experiments, i.e., a net
helicity of 1 was reached at lower ee for 90.
In the study of the inuence of the number of stereogenic
centers on the chiral amplication, the results suggested that
lowering of the number of stereocenters caused an enhancement
of the degree of chiral amplication.224
To quantify the majority-rules and sergeants-and-soldiers data,
Meijer et al. proposed two free energy penalties, i.e., HRP and
MMP, in the chiral amplication in supramolecular assemblies.
Herein, the HRP (helix reversal penalty) describes the energy
penalty of a helix reversal in the aggregate.225,226 This energy
penalty is paid when in a helical stack of these building blocks the
handedness of the stack is reversed, i.e., going from a left-handed
to a right-handed helical segment or vice versa. The HRP value is

asymmetrically substituted monomers have been synthesized to


study the eect of the number of stereocenters and the position
of the stereogenic center on the degree of chiral amplication.
First, when the position of the stereogenic center in asymmetrically substituted BTAs is varied, an oddeven eect can be
discerned, which was characterized by the appearance of a
positive Cotton eect as for (R)-89 and (R)-91 and a negative
Cotton eect for (R)-90. For the (R)-89:88 and (R)-91:88
mixtures, a net helicity of 1 is obtained at a sergeant fraction of
0.15 in both of these systems, while more than 30% sergeant (R)90 is needed to obtain a net helicity of 1 in the (R)-90:88
mixture.
Along with sergeants-and-soldiers experiments, the authors
further performed majority-rules experiments. In contrast with
the weaker amplication of chirality of (R)-90 in sergeants-and7341

DOI: 10.1021/cr500671p
Chem. Rev. 2015, 115, 73047397

Chemical Reviews

Review

Figure 61. CD spectra of mixtures of (R)-89:88 (A), (R)-90:88 (B), and (R)-91:88 (C). Net helicity versus fraction of sergeant for mixtures of (R)89:88, (R)-90:88, and (R)-91:88 (D). Reprinted with permission from ref 224. Copyright 2010 American Chemical Society.

trimer 94 would be able to act as a sergeant and control the


overall helicity of a columnar stack consisting of soldiers of 93.
However, it was surprising to nd that a total absence of
amplication of chirality was found in this system. The author
proposed that the interactions between 93 and 94 were not in the
correct balance to express chirality from the sergeants to the
soldiers. For porphyrin trimers 95 and 96, which, in comparison
to trimers 93 and 94, lack three of the four meso-phenyl rings at
the porphyrin moieties, intermolecular interactions between
these molecules in the columnar assemblies would be stronger
than between the molecules of 93 and 94 in their respective
stacks. This is because 95 and 96 can approach each other as the
result of the reduced steric hindrance and thus enhance the
intermolecular stacking interactions. The amplication of
chirality in toluene was thus very successful, while in n-heptane it
was completely absent. This might be due to the columnar stacks
of porphyrintrimers in n-heptane which are kinetically inert
assemblies, and as a result, the chiral and achiral stacks cannot
dynamically exchange their building blocks. The sergeant and
soldiers experiments have proven to be an excellent method for
revealing this behavior. The work presented here shows the
inuence of variations in the molecular structure and the choice
of solvent amplication of chirality on system chiral
amplication.

related to intermolecular interaction, as once a handedness is


chosen strong intermolecular interactions are favorable to
maintain this handedness throughout the stack. The MMP
(mismatch penalty) is related to the incorporation of a chiral
monomer in a helical aggregate of its unpreferred helicity. In
majority-rules and sergeants-and-soldiers experiments, the HRP
would be very similar, since it is related to the intermolecular
interactions. The MMP, on the other hand, has dierent physical
meaning in the two types of experiments. For the sergeants-andsoldiers experiment, a MMP arises when the chiral sergeant is
incorporated in a stack of achiral molecules of its unpreferred
helicity. For the majority-rules experiment, the MMP arises when
one chiral enantiomer is incorporated in a stack formed from
chiral monomers of opposite stereoconguration with corresponding opposite helicity. For the BTA derivatives discussed
above, the HRP value is similar in all systems, but the MMP is
directly related to the number of stereocenters present in the
molecules. Increasing this number from one to three resulted in
an increase in this energy penalty while leaving the HRP
unaected. These ndings can help gain a better understanding
of the ultimate limits of chiral amplication.
Elemans et al. further illustrated the eect of molecular
structure on chiral amplication through the synthesis of an
analog of porphyrin trimers based on benzenetricarboxyamide
(BTA), 93, 94, 95, and 96.99 In this case, the chiral porphyrin
7342

DOI: 10.1021/cr500671p
Chem. Rev. 2015, 115, 73047397

Chemical Reviews

Review

Figure 62. (A) Molecular structures of the porphyrin trimers 9396. (B) Schematic representation of the self-assembly of porphyrin trimers in helical
columnar stacks. Reprinted with permission from ref 99. Copyright 2012 Royal Society of Chemistry.

4.4.2. Chiral Amplication in Binary Systems. Chiral


amplication was also studied in binary complex systems.227 In
this case, melamines equipped with two PBI chromophores and
two 3,7-dimethyloctyl chiral handles were mixed with cyanuric
acid to form a discotic supramolecular complex (CA).228 It was
found that the sergeants-and-soldiers eect where a few chiral
building blocks can control the helical sense of the large number
of structurally related achiral ones was not applicable for this
system. When mixed chiral 97S was used as sergeant with the
optically inactive achiral 97A as the soldiers, plots of versus
the amount of 97S for these ternary mixtures (97S/97A/CA)
showed a decrease in their optical activities compared to the
chiral complex, indicating the absence of any sergeant-andsoldiers eect. In contrast to the sergeants-and-soldiers eect, the
other chiral amplication eect, the majority-rules eect,
occurred at the level of the hydrogen-bonded complexes.
Enantiomeric mixtures of 97S and 97R showed almost linear
dependence of on ee in the absence of CA (recipe i),
suggesting that the self-aggregation of the enantiomers (selfsorting) or coaggregation occurs; they obey their own preferred
helicities. In the presence of CA (recipe ii), the /ee plots

indeed diverged from linearity, indicating that the amplication


of chirality obeyed the majority rules and occurred at the level of
the hydrogen-bonded complexes.
4.4.3. Chiral Amplication to Nanoscale. Chiral
amplication can be expressed in the form of nano/microstructures. Compound 98 is an achiral molecule and the analogue
of compound 99, which has four chiral centers in its alkyl chains.
The self-assembly of 98 could only give a at nanostructure.
However, when 98 was coassembled with 99, the at lamellae
were transformed into twisted ribbons.229 The presence of only 5
mol % of the sergeant 99 was able to transfer the chirality
embedded in the peripheral chains to the remaining 95% of the
soldiers (98), as revealed by the corresponding SEM images, in
which the micrometer-long twisted ribbons of high aspect ratio
appeared. The presence of stereogenic centers in the coassembly
of achiral 98 with chiral 99 provoked the chiral propagation of
the H bonding of the amide functionalities reinforcing the
formation of twisted ribbons. Increasing the percentage of chiral
sergeant 99 in the coassembly caused the formation of twisted
ribbons to increase. The results represent an excellent example of
the study of homochirality on surfaces and, at the same time,
7343

DOI: 10.1021/cr500671p
Chem. Rev. 2015, 115, 73047397

Chemical Reviews

Review

Figure 63. (a) Chemical structures of 97 and CA. (b) Proposed structure of 3:1 hydrogen-bonded complex 97SCA. (c) Schematic illustration of helical
columns. Reprinted with permission from ref 228. Copyright 2011 John Wiley & Sons.

4.5. Chiral Memory in Supramolecular Systems

contribute to the knowledge about the set of rules governing the


generation of chiral objects that hold great potential for the
development of supramolecular devices.
4.4.4. Unexpected Amplication in Racemate Assemblies. Liu et al. found an interesting chiral amplication in selfassembled systems based on L- and D-alanine derivatives
containing an N-uorenyl-9-methoxycarbonyl (Fmoc) moiety
and a long alkyl chain.230 It has been found that both the
enantiomeric and the racemic assemblies showed CD signals.
The enantiomer 100 showed mirror-imaged CD spectra, and the
sign of the CD spectra for the assemblies followed the molecular
chirality. However, the supramolecular chirality of the racemate
assembly was not certain. It was revealed that the slight excess of
one enantiomer in the racemic mixtures may result in an active
CD signal since the exact 1:1 mixture at a molecular level cannot
be reached. When mixing the two enantiomers (100L/100D)
with dierent molar ratios it was found that an excess of 100L
resulted in a negative Cotton eect, whereas mirror-imaged CD
spectra were obtained for mixtures with an excess of 100D.
Furthermore, the CD signals observed for a nonequimolar
mixture of the enantiomers were more intense than those for the
pure enantiomers. For the system obeying the majority-rules
principle, the CD intensity generally decreased when their
mixing ratio deviated from the pure one. However, in the case of
the mixed 100L/100D system, the CD signals intensied when
the mixing ratio approached 1:1, as shown in Figure 65. This
indicated that the self-assembly of the racemic mixture is very
sensitive to a slight enantiomeric excess and the system could be
used for the detection of a broad range of chiral amino acid
derivatives.

The phenomenon of chirality memory describes a supramolecular system in which chirality is rst induced and then
maintained after the chiral source is erased or replaced by an
achiral component. Thus, these complexes have a memory for
the chirality of the species that induced the systems asymmetry
after the removal of these inductor molecules.
Generally, chiral memory is dicult to induce in noncovalent
supramolecular assemblies, partly because additives often
interfere with the noncovalent interactions that hold the
assembly together. However, in recent decades, more eorts to
explore supramolecular chirality memory systems have proven to
be successful through the eorts to design the chiral and achiral
units and control the dynamics. In order to realize the chiral
memory, there are several important elements. First, the induced
chiral nanostructure should be generally stable. Thus, even when
you remove the chiral species, the chirality is maintained. Second,
a small amount of the chiral substance should be able to induce
chirality in the system. A successful chiral memory system may
contain (1) noncovalently induced helical polymers, (2) strong
aggregates from achiral building blocks such as J or H aggregates,
or (3) chiral cages from coordination compounds.
4.5.1. Helicity Memory in Noncovalently-Induced
Helical Polymers. Helicity memory in the noncovalently
induced helical polymers has been thoroughly investigated by
Yashima and co-workers.174 On the basis of a chiral memory
system of noncovalent helicity induction in optically active
polymers, Yashima et al. developed an excellent macromolecular
memory system of a helical polyacetylene in the solid state.231
They synthesized a polyacetylene derivative, 101, and induced its
preferred-handed helicity in the presence of (S)-phenylethanol in
n-hexane through weak hydrogen-bonding interactions. 101 was
7344

DOI: 10.1021/cr500671p
Chem. Rev. 2015, 115, 73047397

Chemical Reviews

Review

Figure 65. (A) G values centered at 309 and 255 nm as a function of the
ee value of nonequimolar mixtures of 100L and 100D. The G value of
the racemate was set to zero. (B) SEM images of the nanostructures
formed from mixtures of 100L and 100D with various ee values.
Reprinted with permission from ref 230. Copyright 2013 John Wiley &
Sons.

Figure 64. (Top) Structure of bisamides 98 and 99, schematic


illustration of the self-assembly of 98 into sheets, and the coassembly of
98 and 99 into twisted ribbons. (Bottom) SEM images of twisted
ribbons formed by the coassembly of the achiral soldier 98 and chiral
sergeant 99 at dierent percentages of chiral component: 95/5 (a and
b), 90/10 (c), and 85/15 (d). Reprinted with permission from ref 229.
Copyright 2010 Royal Society of Chemistry.

then recovered by ltration followed by washing with methanol


to completely remove (S)-phenylethanol. Upon further
dissolving 101 in n-hexane at 20 C, an apparent ICD band
exhibited an intensity that increased with time. This intensity
nally became nearly equal to that induced in an n-hexane
solution in the presence of (S)-phenylethanol after 1 h. The
macromolecular helicity memory in the solid state was more
stable than in solution, because the ICD intensity in the solid
state persisted for at least 11 months at 25 C. More interestingly,
by simply immersing the polymer with macromolecular helicity
memory by (S)-phenylethanol in a solution containing (R)phenylethanol followed by washing with methanol, the helicity of
101 was completely inverted and memorized. Thus, there was no
doubt that a preferred-handed helix of the optically inactive
polymer 101 was induced via weak noncovalent bonding
interactions in the solid state simply by immersing 101 in a
nonracemic liquid, and this helicity could be memorized and
switched automatically. This chiral memory and switchable
system in the solid state will be benecial to the development of a
chiral stationary phase for the separation of enantiomers.
Inoue and Takashima et al. designed a m-ethynylpyridine
polymer that has a metal coordination site at the 4 position of
each pyridine unit, which showed a chiral memory eect on a methynylpyridine oligomer.232 It is found that the polymer can
form CD-active helical complexes with various kinds of guest
saccharides by the interaction of hydrogen bonds between the

Figure 66. Reversible switching and memory of macromolecular helicity


of 101 in the solid state. Preferred-handed macromolecular helicity of
101 is induced and subsequently memorized in the optically inactive
101 via noncovalent interactions with a nonracemic alcohol (S- or Rphenylethanol) followed by complete removal of phenylethanol in the
solid state. The polymers helical handedness and axial twist sense are
switched reversibly in the solid state in the presence of the opposite
enantiomeric alcohol. Reprinted with permission from ref 231.
Copyright 2014 Nature Publishing Group.
7345

DOI: 10.1021/cr500671p
Chem. Rev. 2015, 115, 73047397

Chemical Reviews

Review

nitrogen atoms of the pyridine rings and the hydroxy groups of


the saccharides. Moreover, the ICD band was remarkably
enhanced by the addition of Cu(OTf)2 (0.5 equiv to pyridine
units) and o-phenanthroline (phen) as a result of stabilization of
the helical structure of the polymer. Even when an equimolar
amount of another enantiomer of glucopyranoside was added to
make the whole system apparently racemic, the ICD of the
ethynylpyridine polymer was memorized and remained for
several weeks.

Figure 67. Assumed mechanism of helix stabilization of methynylpyridine host oligomer 102 by coordination of copper and
phen inside the helix. Reprinted with permission from ref 232.
Copyright 2012 Royal Society of Chemistry.

4.5.2. Chiral Memory in Aggregates Such as J and H


Aggregates. Purrello et al. reported that a number of selfassembly systems could memorize the chirality of complexes
when the chiral auxiliary had been removed. They rst induced
the formation of aggregates between the cationic porphyrin
CuTMP and anionic porphyrin H4TPPS, 63, with the chiral
matrix of either L- or D-polyglutamic acid.233 Upon the formation
of the ternary complex, the induced CD of binary CuTMP and
TPPS was obtained in the -helix structure of polyglutamic acid.
Interestingly, the induced CD signals of the porphyrin complexes
remained even when the conformation of the polyglutamic acid
was switched to a random coil by increasing the pH to 12. This
suggested that during the formation of the ternary complex the
chirality of the polymer was transferred to and memorized by the
porphyrin complexes. Interestingly, the chiral complex was stable
and maintained even when a 5-fold excess of a competing and
antipodal chirality source; namely, poly-D-glutamate, was added
to the CuTMP/TPPS/poly-L-glutamate system. A similar selfassembly memory system based on the aggregation of the
oppositely charged CuTMP and TPPS in the presence of
enantiopure aromatic amino acids was also found.234238
Further, Shi et al. and He et al. found simple TPPS J aggregates
also showed chiral memory eects.239242 For instance, He et
al.242 found that with the existence of L- or D-enantiomers of cis[CoBr(NH3)(en)2]Br2 as chiral triggers for the J aggregates of
achiral 5,10,15,20-tetrakis(4-sulfonatophenyl)porphyrin
(TPPS), 67 could be fabricated into chiral assemblies, during
which the metal-centered chirality can be transferred to the J
aggregates. In addition, the chirality was memorized in the
porphyrin J aggregates. These authors initially synthesized the
porphyrin aggregates with the L-Co(III) complex, and then the DCo(III) complex was added to the system. During the addition of
this opposite D-Co(III) chiral species, the UV absorption at the
TPPS Soret band (434 nm) gradually decreased; however, the
induced CD signal of the porphyrin aggregates increased (rather
than inverting) after the addition of excess D-Co(III) complex.
This suggested that the chirality of TPPS J aggregates induced by
one enantiomer of Co(III) coordination was maintained even
with the addition of an excess of the opposite enantiomer of the
Co(III) complex.
Meijer et al. developed a class of highly tunable porphyrinbased chiral memory system based on the chiral Zn porphyrins

Figure 68. Schematic structures of H2TPPS and - or -Co(III)


complexes and Schematic illustration of the induction, memory, and
amplication of chirality in H2TPPS with - or -Co(III) complexes.
Reprinted with permission from ref 242. Copyright 2010 Royal Society
of Chemistry.

and achiral Cu porphyrins through a dynamic control of


assemblies.243,244 The chiral Zn porphyrins were used as a
sergeant to transfer their chirality to the achiral Cuporphyrin
soldiers. After the sergeant was removed from the coaggregates
by axial ligation with a Lewis base (quinuclidine), the chiral
information in the remaining aggregate was preserved as a result
of slow conformational dynamics, which revealed a chiral
memory eect.
Such chiral memory based on the aggregation has been
expanded to other systems. Jiang et al. established a J aggregate of
an achiral perylene dianhydride (PDA) in CTAB micelle solution
and employed small molecule D- and/or L-tartaric acid as the
chiral auxiliary to induce transfer of the chirality to PDA
aggregates.245 The ICD signal of the J aggregates was also found
to remain unchanged upon addition of a large excess of the
alternative enantiomer of tartaric acid, implying the imprinting of
the chirality in the J aggregates.
Aida et al. synthesized an elaborate nanotubular helical
architecture with 60% de (80:20 diastereomeric ratio) by the selfassembly of a hexabenzocoronene derivative, HBCPy, carrying a
chiral (BINAP)Pt(II) moiety as a detachable chiral auxiliary. The
optically active nanotubes did not racemize after removing the
chiral auxiliary through the addition of ethylenediamine. Once
the helical tubular structure of HBC formed, the addition of
(BINAP)Pt(II) with an absolute conguration opposite to the
original one did not cause the helical inversion. These results
7346

DOI: 10.1021/cr500671p
Chem. Rev. 2015, 115, 73047397

Chemical Reviews

Review

stereocenters of the structure enabled retention of conguration


upon replacement of the chiral subcomponents. This memory
eect allows for the stereoselective preparation of a metal
organic capsule that ultimately contains only achiral subcomponents, which can extend the application of metalorganic
capsules in stereoselective guest recognition and sensing and as
asymmetric reaction vessels.

5. SPONTANEOUS SYMMETRY BREAKING AND


EMERGENCE OF SUPRAMOLECULAR CHIRALITY IN
SELF-ASSEMBLED SYSTEMS FROM EXCLUSIVELY
ACHIRAL MOLECULES
Through self-assembly, not only chiral molecules but also
completely achiral molecules can form chiral supramolecular
assemblies. This situation results from spontaneous symmetry
breaking,10 which is one of the most important issues in
obtaining assemblies with macroscopic chirality248 or optical
activity249 instead of producing the same amount of the
enantiomeric nanostructures. On the other hand, there is a real
possibility that the origin of life could have depended on
molecular chirality and supramolecular chirality.250 Although we
still do not clearly know the origin of natural homochirality,
supramolecular chirality can now be created from achiral
molecular building blocks. In general, asymmetric environments
are necessary for symmetry breaking, which provide the
supramolecular chirality to the assemblies of achiral molecules.236,251253 However, the most intriguing possibility would
be symmetry breaking without discernible chiral conditions. In
this section, we will show examples where supramolecular
chirality emerged from achiral building blocks.

Figure 69. (A) Chiral/achiral amide-functionalized zinc/copper


tetraphenylporphyrins. (B) Schematic depiction of selective depolymerization with chirality retention and temperature-induced switching of
the chiral memory. Reprinted with permission from ref 244. Copyright
2010 American Chemical Society.

demonstrated a good example of a stereochemical memory eect


in HBC nanotubular helical architecture.246
4.5.3. Helicity Memory in Chiral Cages from Coordination Compounds. A class of FeII4L4 capsules with chirotopic
cavities has been prepared by the in situ metal-templated imine
condensation between tris(formylpyridyl) benzene and a chiral
amine.247 This capsule maintained the stereochemistry of the
cage framework (99% ee) even when the chiral amine was
replaced by an achiral one. The cage retained its stereochemistry
after 4 days at 90 C. The author inferred that the strong
cooperative stereochemical coupling between the iron(II)

Figure 70. Schematic illustration of a series of experiments for investigating the dynamic nature of nanotubularly assembled HBCPy. Reprinted with
permission from ref 246. Copyright 2013 American Chemical Society.
7347

DOI: 10.1021/cr500671p
Chem. Rev. 2015, 115, 73047397

Chemical Reviews

Review

Figure 71. Route i: formation of racemic cage 2 through subcomponent


self-assembly. Route ii then route iii: enantioselective formation of cage
2 through subcomponent substitution. Reprinted with permission from
ref 247. Copyright 2013 American Chemical Society.

5.1. Liquid-Crystal and Banana-Shaped Molecules

Even though the origin of chirality and natural homochirality has


attracted much attention over the past decades, supramolecular
chirality resulting from the self-assembly of achiral molecular
building blocks has been of major interest as a result of the early
work conducted on liquid crystals. The rst report of this can be
found in the work of Young and co-workers.254 These authors
designed and synthesized a series of stilbene derivatives and
studied the possibility of forming nematic liquid crystals from the
assembly of these molecules. Interestingly, with the racemic
stilbene mixtures they observed small cholesteric liquidcrystalline regions in the nematic eld under polarized light
microscopy.
In the use of achiral molecules to form chiral liquid crystals, the
most famous building blocks are the banana-shaped or bent-core
achiral molecules.255 In the eld of liquid crystals, the selfassembly of banana-shaped achiral molecules has been widely
investigated. For example, Tschierske and colleagues studied the
liquid-crystalline phases formed by silicon-containing polyphilic
bent-core achiral molecules 104. These authors found that
temperature-induced inversion of chirality in a supramolecular
system formed by achiral molecules can be established (Figure
72).256
The research group of Cheng constructed chiral propellers
from the self-assembly of achiral molecules (BPCA-Cn-PmOH),
which were composed of 4-biphenylcarboxylic acids connected
with phenol via alkoxyl chains 105. The achiral BPCA-CnPmOH molecules can form individual head-to-head dimers, and
the twisting of the dimers can lead to chiral N phases. These
results demonstrate that neither molecular chirality nor a
molecular bend is necessary to form a chiral phase (Figure 73).257
For chiral assemblies formed by banana-shaped (bent-core)
achiral molecules, a notable work is that of Hough, Clark, and co-

Figure 72. Molecular structures of silicon-containing bent-core achiral


molecules, and the temperature-induced inversion of supramolecular
chirality. Reprinted with permission from ref 256. Copyright 2006
American Chemical Society.

workers (Figure 74).258 In this instance, the self-assembly of


banana-shaped (bent-core) achiral molecules 106 can result in
strong supramolecular chirality. Most interestingly, these
assemblies do not exhibit anisotropy at the macroscopic scale.
These assemblies are macroscopically isotropic uids that
possess only short-range orientation and positional order, just
like a true liquid. Therefore, the self-assembly of achiral
molecules was found to form isotropic uids with supramolecular
chirality. In particular, the assembly of bent-core achiral
molecules can form a phase that exhibits smectic layering with
uid order within the smectic layers. Thus, the coherent length of
the smectic layering is very short, smaller than ca. 100 nm. The
mechanism of formation of this property has been attributed to
the formation of saddlesplay deformations involving the elastic
constant K24. Saddlesplay director elds are not space lling,
and smectic layers are unstable in saddlesplay deformations and
thus incompatible with long-range order.
With respect to the banana-shaped (bent-core) achiral
molecules forming supramolecular chirality, Hough, Clark, and
co-workers (Figure 75) also studied helical nanolament phases,
in which a local chiral structure is expressed as twisted layers.259
Although composed of achiral molecules 107, the layers in these
laments are twisted and rigorously homochiral, demonstrating
broken symmetry. This work was published in the same issue of
Science as that of isotropic chiral uids. In contrast to isotropic
7348

DOI: 10.1021/cr500671p
Chem. Rev. 2015, 115, 73047397

Chemical Reviews

Review

Figure 73. (A) Structure of the dimeric building block of BPCA-C7PmOH 105. (B) Chiral propellers from the self-assembly of this achiral
molecule 105 (BPCA-Cn-PmOH). Reprinted with permission from ref
257. Copyright 2006 John Wiley & Sons.
Figure 75. (A) Structure and phase sequence upon cooling of the four
B4 phase-forming compounds studied. (B) Mechanism and TEM
images of hierarchical self-assembly of the nanolament (NF) phase
starting with bent-core mesogenic molecules. Reprinted with permission
from ref 259. Copyright 2009 The American Association for the
Advancement of Science.

both bent-core molecules and rod-like molecules.260 This


process can be considered a general method for producing
homochiral helical nanolaments. Banana-shaped molecules
(108, 109) are in the B4 phase, while rod-like molecules (110,
111) can form a nematic phase. The mixture of dierent
molecules has the phase sequence NBx(B4/N), and homochiral helical nanolaments can be obtained upon cooling
(Figure 76).261
To understand the helical nanolament phases assembled by
banana-shaped (bent-core) achiral molecules, the corresponding
hierarchical nanostructures have attracted much attention. It was
found that the arrangement of dierent helical nanolaments can
be very important to the assembly process. Jakli et al. found that
some properties, especially structural color, of the B4-phase
liquid crystals formed by bent-core achiral molecules 112 cannot
be explained by the nanostructures of nanolaments alone. In
their study, they found that dierent helical nanolaments do not
form parallel packing. Instead, these helical nanolaments were
arranged at an angle of 3540 with respect to each other,
forming a doubly twisted nanostructure, which caused the
unusual structural color of these liquid crystals (Figure 77).262

Figure 74. (A) Structures, phase sequences, layer spacing, and


correlation lengths (as measured by XRD and freezefracture TEM)
of achiral 106. Freezefracture TEM images of the dc phase of 106. (B)
Frustration between molecular fragments in a tilted bent-core molecule
can be relieved by saddle-splay curvature of the layers. Reprinted with
permission from ref 258. Copyright 2009 The American Association for
the Advancement of Science.

5.2. Solution Systems, Micelles

Asymmetric environments or conditions are usually necessary for


symmetry breaking within supramolecular systems, in which the
aggregation of achiral molecular building blocks can lead to
supramolecular chiral assemblies. Thus, chiral dopants202 or a
chiral matrix263 can assist achiral molecules to assemble into
chiral nanostructures. Solid surfaces264,265 also can provide twodimensional conned environments that lead to the symmetry
breaking.

chiral uids assembled by achiral building blocks, these helical


nanolament phases exhibit birefringence, indicating long-scale
ordering.
In the production of helical nanolament phases from the
assembly of banana-shaped (bent-core) achiral molecules,
Takezoe and co-workers developed a mixture system containing
7349

DOI: 10.1021/cr500671p
Chem. Rev. 2015, 115, 73047397

Chemical Reviews

Review

obvious approaches to make the achiral molecules form chiral


assemblies in organic or aqueous solution. Indeed, there are some
reports showing that achiral molecules self-assemble into helical
nanostructures. However, within these systems, the mixture of
both left- and right-handed helices in equal amounts does not
fulll the requirements for true supramolecular chirality.266268
In this case, the assemblies are overall racemic mixtures and show
no macroscopic optical activity with a silent CD.
Researchers in the eld of supramolecular chirality are very
interested in the formation of assemblies with unequal amounts
of left- and right-handed helices using purely achiral molecular
building blocks in solution. This type of symmetry breaking
should be able to be detected by CD spectral measurements.
There are only a few published reports of the aggregation of
achiral molecules in solution to form supramolecular chirality.
Amazingly, the rst paper on this issue was published in 1996.269
Dahne and co-workers found that one achiral charged dye
molecule containing long alkyl chains (113) can form supramolecular chiral assemblies in solution. In this system, two types
of noncovalent interactions play very important roles. Thus, both
the self-association of organic dyes and hydrophobic interactions
between long alkyl chains cause the achiral molecules to undergo
J aggregation with supramolecular chirality, which can be
conrmed by CD spectra. It is suggested that the dye molecule
can form twisted herringbone-type assemblies, which may
possibly lead to supramolecular chirality (Figure 78).
The Liu group studied the aggregation of achiral molecular
building blocks for forming chiral assemblies in oil/water mixed
micelle dispersions.270 The work is based on a surfactant-assisted
self-assembly (SAS). In this work, an aqueous solution of
cetyltrimethylammonium bromide (CTAB) and a chloroform
solution of another achiral molecular building block was added

Figure 76. (A) Chemical structures of bent-core and rod-like


components of the mixtures: (a) 12OAzo5AzoO12 (108), (b) P8-OPIMB (109), (c) 5CB (110), and (d) 5PCB (111). (B) AFM image of
Bx surface. (C) (a) UVvis absorption spectra of 12OAzo5AzoO12
(108) (bent-core) and ZLI2293 (rod-like). (b) CD spectra of the Bx
phase. (c) CD spectra showing a longer wavelength peak made from TN
cells of various cell thicknesses: 0.4, 0.6, 0.7, and 0.9 m. (Inset) CD
intensity as a function of cell thickness. Reprinted with permission from
ref 261. Copyright 2013 John Wiley & Sons.

In a homogeneous solution, the birth of chiral information


from totally achiral systems can be very subtle, so there are no

Figure 77. (A) Molecular structure and phase sequence of PnOPIMB (112) (n = 7, 8, 9 and 12) during cooling at 1 C min1. (B) TEM image and
models showing the double-twist structure. Reprinted with permission from ref 262. Copyright 2014 Nature Publishing Group.
7350

DOI: 10.1021/cr500671p
Chem. Rev. 2015, 115, 73047397

Chemical Reviews

Review

assembly of ZnTPyP. Interestingly, no supramolecular chirality


was detected in these nanotubes or nanobers (Figure 79).
With regard to symmetry breaking upon self-assembly,
dierent achiral porphyrins have been used as model molecular
building blocks. In aqueous solution, the self-assembly of some
water-soluble porphyrins (such as tetrakis(4-sulfonatophenyl)porphine, TPPS, 67) has been widely investigated and is further
discussed below. On the other hand, the uncharged achiral watersoluble porphyrin was also found to form chiral assemblies in
aqueous solution.
Mineo et al. synthesized a TPP-based porphyrin containing
polyethyleneoxy substituents (PPeg4, 115).271 This uncharged
porphyrin can spontaneously form self-assembled structures in
water. Although PPeg4 is achiral, this molecule forms a chiral
assembly in water. Most interestingly, the supramolecular
chirality from the assembly of PPeg4 was found to be induced
by a weak thermal force. In this case, the intensity of the CD
signal of the PPeg4 assemblies could be enhanced by increasing
the temperature (Figure 80).
Meijer and co-workers recently investigated the symmetry
breaking of an assembly of achiral molecular building blocks in
organic solvents.272 They found that the self-assembly of achiral
partially uorinated benzene-1,3,5-tricarboxamide molecules
116 could produce supramolecular chirality (Figure 81A). For
these systems, they carefully studied the kinetics of the selfassembly, and the unique two-step self-assembly behavior was
conrmed. Moreover, they found that true symmetry breaking
could happen during a kinetically controlled secondary
nucleation (Figure 81B). Thus, within the initial self-assembly
process, achiral partially uorinated benzene-1,3,5-tricarboxamide molecules could assemble into equal amounts of onedimensional left- (M) and right-handed (P) helical aggregates,
whereas the systems as a whole did not show any optical activity.
During kinetically controlled secondary nucleation, these onedimensional helical aggregates could bundle into nanobers,
which could be optically active.272 Here, the partially uorinated
molecular structures are important, which produced the chiral
bias that was then amplied by dierent noncovalent
interactions. Hydrogen bonding was considered to dominate

Figure 78. Spontaneous formation of supramolecular chirality in J


aggregates.

dropwise into the aqueous solution with stirring. This method is


related to the formation of a microemulsion, where the
chloroform evaporates during the self-assembly process. Using
the SAS method via an oil/aqueous medium, supramolecular
assemblies with dierent nanostructures and properties could be
obtained in a controllable manner from very simple achiral
molecular building blocks. For example, when a chloroform
solution of zinc 5,10,15,20-tetra(4-pyridyl)-21H,23H-porphine
(ZnTPyP) (114) was added dropwise to a cetyltrimethylammonium bromide (CTAB) aqueous solution, dierent nanostructures with varying supramolecular chirality can be obtained
depending on the concentration of CTAB and the aging time.
Using the 0.9 mM CTAB system for assembly with subsequent
aging for 3 days, chiral nanorods can be produced by the CTABassisted self-assembly of ZnTPyP. The supramolecular chirality
was proven by CD spectral measurements. However, when the
concentration of CTAB and the time for aging were changed,
dierent supramolecular nanostructures, such as nanotubes and
long nanobers, were produced by the CTAB-assisted self-

Figure 79. Schematic illustration showing the controlled synthesis of various porphyrin nanostructures with varied supramolecular chirality by means of
an SAS, where an oil/aqueous medium was employed. CD spectra of ZnTPyP (114) nanorods fabricated in sample B that was aged for 3 days. Black and
red curves are the results detected from the samples prepared in dierent batches. Reprinted with permission from ref 270. Copyright 2010 American
Chemical Society.
7351

DOI: 10.1021/cr500671p
Chem. Rev. 2015, 115, 73047397

Chemical Reviews

Review

Figure 80. (A) Structure of PPeg4 (115). (B) Circular dichroism


spectra at T = 26 C (full line), 30 C (dotted line), and 36 C (dashed
line) of the PPeg4 aqueous solution. Reprinted with permission from ref
271. Copyright 2014 The Royal Society of Chemistry.
Figure 81. (A) Chemical structures of the partially uorinated BTAs
(116). (B) (a) AFM image of drop-casted solutions of BTA-F8H on
mica (scale bar = 0.5 mm). (b and c) Results of stopped-ow
experiments for solutions of BTA-F8H. (d) Self-assembly mechanism of
116. Reprinted with permission from ref 272. Copyright 2012 John
Wiley & Sons.

the formation of the helical aggregates, while dipoledipole


interactions dominate the secondary nucleation.
Hao and Sun et al. found that achiral bolaamphiphilic
azobenzene 117 was able to form chiral assemblies in aqueous
solution.199 It was shown that the precipitation process was
important. The achiral bolaamphiphilic azobenzene containing
many carboxyl groups (Figure 82A) was found to dissolve in
water at pH 6.56, but when the pH value was adjusted to 2.77 by
adding HCl solution, the bolaamphiphilic azobenzene formed
precipitates. TEM and SEM measurements of the material
showed that these precipitates consisted of nanotubes with
supramolecular chirality (Figure 82B), which was detected by
CD spectral measurements (Figure 82C). Furthermore, since the
azobenzene derivatives are photosensitive, the morphology and
chirality of the self-assembled nanostructures derived from the
achiral bolaamphiphilic azobenzene could change in response to
external stimuli such as light and heat.
An acidication process was also implemented for the selfassembly of TPPS (67) in aqueous solution. TPPS is a wellknown building block for the occurrence of symmetry breaking
during the self-assembly. The acidication of TPPS could help in
the formation of J aggregates with supramolecular chirality.
Interestingly, the emergence of supramolecular chirality on
TPPS assemblies in aqueous solution can be controlled
kinetically by modulating the speed of the acidication of the
porphyrin. In a recently published paper by Scolaro et al., detailed
kinetic investigations of the self-assembly of acidied TPPS
demonstrate that the rate of the aggregation process strongly
aects the chiral induction.273 In this study, the aggregation
process in aqueous solution could be changed by adding the
porphyrin as the rst (PF) or last reagent (PL). PF represents the
slow aggregation process, while PL represents the fast acidication and aggregation process. The CD spectral measure-

ments show that only the slow aggregation process can lead to
supramolecular chiral assembly (Figure 83).
For symmetry breaking upon the self-assembly of achiral
building blocks, a recently published THF/water system is worth
mentioning. Since THF and water are not completely miscible, a
liquid/liquid interface is present in THF/water mixtures with
volume ratios of 1/1. When a THF solution of achiral carboxyl
azobenzene derivatives (Figure 84A) was added dropwise to an
aqueous solution of melamine, supramolecular chiral assemblies
can be obtained at the THF/water interface along with the
permeation and volatilization of THF (Figure 84C).274 For the
formation of coassemblies, hydrogen bonding between carboxyl
azobenzene derivatives and melamine plays a very important
role. TEM and AFM measurements show that these assemblies
are long and helical bers with intrinsic conformational chirality
(Figure 84B). Furthermore, the morphology and chirality of the
supramolecular assemblies are photoresponsive, which is
induced by the photoisomerization of the azobenzene
components within the self-assembled nanostructures.
For supramolecular chirality from an assembly of achiral
molecules in solution, it seems that -conjugated molecules are
always necessary. Within these self-assemblies, the aromatic rings
tend to overlap through stacking. Thus, the displacement of
aromatic rings within the molecular packing shows a slight angle
between neighboring rings, which will produce a chiral bias. If
this bias is repeated by forming a helix in a certain direction,
supramolecular chirality would emerge and be amplied.
7352

DOI: 10.1021/cr500671p
Chem. Rev. 2015, 115, 73047397

Chemical Reviews

Review

Figure 82. (A) Molecular structure of achiral bolaamphiphilic azobenzene 117 and its packing mode within the self-assembly. (B) (a) TEM and (b)
SEM observation of the self-assembled coiled and tubular nanostructures of 117 from water at pH 2.77. (c) TEM image of the twisted ribbon. (d) HRTEM image of the nanotube. (C) CD spectra of the self-assembled suspension 117 at pH 2.77 (solid line) and 117 itself (dashed line). Reprinted with
permission from ref 199. Copyright 2012 The Royal Society of Chemistry.

Figure 83. (A) Molecular structure of TPPS (67). (B) Schematic illustration showing kinetic control of the supramolecular chirality of the assembly of
acidied TPPS with J aggregation. Reprinted with permission from ref 273. Copyright 2014 American Chemical Society.

Figure 84. (A) Chemical structures of carboxyl azobenzene derivatives and melamine, and the schematic representation of their complex. (B) (a, b)
TEM and (c and d) AFM observation of the self-assembled helixes and supercoils with labeled handedness. (C) CD measurements of the assemblies.
Reprinted with permission from ref 274. Copyright 2014 The Royal Society of Chemistry.

5.3. Gel Systems

helps the formation of hydrogen bonding. Thus, spontaneous


chiral symmetry breaking through the steric eect of imidazole
upon gelation was achieved.282
Another chiral supramolecular gel assembled by achiral
molecules is also related to the complexation of the imidazole
unit with metal ions. You and co-workers demonstrated that
simple achiral molecules containing an imidazole unit could form
supramolecular polymers by complexation with Ag+, which can
further gel in a variety of solvents. The supramolecular metal gels
formed by simple achiral molecules can be optically active.283
Recently, Liu and Wang et al. synthesized an achiral C3symmetric benzene-1,3,5-tricarboxamide substituted with ethyl
cinnamate (BTAC, 119) and studied its supramolecular gelation
and macroscopic chirality from the self-assembly of BTAC in the
DMF/H2O mixture. They found that upon gelation this achiral
compound can simultaneously self-assemble into unequal
amounts of left- and right-handed twists, thus resulting in
macroscopic chirality without any chiral additives (Figure 86).
The symmetry breaking and formation of macroscopic chirality
from the assembly of this type of molecules is quite rare. The
hierarchical self-assembly of an uneven number of dierent chiral
assemblies produces the unbalanced left- and right-handed

As a very important form of soft matter with extensive potential


applications, supramolecular gels fabricated by noncovalent
bonds have attracted great interest recently.275280 Chiral
supramolecular gels are usually fabricated by chiral gelators and
show macroscopic optical activity and/or helical nanostructures.
A few achiral gelators have been found to self-assemble into
helical nanostructures with an equivalent amount of left- and
right-handedness,266,268,281 but macroscopic optical activity was
barely detected from these supramolecular gels. However, there
are really some examples of optically active gels obtained from
achiral gelators.
Kimura et al. synthesized an achiral disk-shaped molecule
having one imidazole unit (Figure 85A). This achiral molecule
118 was found to form a supramolecular gel in 2-methoxyethanol
and assemble into long and twisted nanoscopic bers (Figure
85C). Although the molecular building block is achiral, it forms
supramolecular gels with optical activity, which can be
demonstrated by CD spectral measurements (Figure 85B).
During the emergence of the optical activity, it was found that the
imidazole substituent is very important, because it increases the
asymmetrical characteristics of the molecular building block and
7353

DOI: 10.1021/cr500671p
Chem. Rev. 2015, 115, 73047397

Chemical Reviews

Review

LangmuirBlodgett (LB) lms through the air/water interfacial


self-assembly.
The Liu group rst reported air/water interfacial chiral
assemblies from achiral molecular building blocks in 2003.39 In
this work, it was found that when an achiral amphiphilic molecule
containing a naphtha[2,3]imidazole headgroup and long alkyl
chain (NpImC17, 120) was assembled with Ag(I) ions at the air/
water interface, the compound could coordinate with AgNO3 to
form a stable monolayer, which can be then further transferred
onto a solid substrate to produce a LangmuirBlodgett (LB)
lm. The CD spectra of these LB lms showed a strong Cotton
eect after removing the possible LD eects, which suggested
that the supramolecular chirality resulted from the assembly of
achiral molecular building blocks. It was suggested that the
overcrowded stacking of the achiral chromophores in a helical
sense would produce such macroscopic supramolecular
chirality.40 Interestingly, in this system containing only achiral
molecules, the sign of the supramolecular chirality cannot be
determined in the air/water interfacial assembly. One can obtain
M- or P-chiral assemblies in dierent batches. It is worth
mentioning that air/water interfacial chiral assemblies obtained
from achiral molecules are totally dierent from those obtained
from racemic mixtures, in which no optical activity can be
detected. In contrast, these LB lms are optically active, even
though the handedness of the supramolecular chirality can
change from batch to batch.
With air/water interfacial assembly, it is not rare to obtain the
supramolecular chirality from achiral molecular building blocks.
For many achiral molecules, this process of making supramolecular chiral assemblies is feasible. For example, Liu et al.
further designed an achiral amphiphilic barbituric acid 121 and
obtained supramolecular chiral assemblies with optical activity at
the air/water interface (Figure 88A). These molecules further
self-assembled into spiral nanoarchitectures (Figure 88B). The
hydrogen bonding between the headgroups and the overcrowded arrangement of the aromatic ring within the assembly
played very important roles.285
The relationship between molecular structures of achiral
molecules and the supramolecular chirality of their assemblies at
the air/water interface has been thoroughly investigated. It was
found that many achiral molecules with larger steric hindrance
during self-assembly can form chiral assemblies on the surface of
water. For example, some achiral arylbenzimidazoles with 2substituted anthryl groups were found to interact with AgNO3 in
the subphase and form chiral assemblies.286 In the case of two
achiral coumarin derivatives, 7-octadecyloxylcoumarin (7CUMC18, 122) and 4-octadecyloxylcoumarin (4-CUMC18,
123) substituted at dierent positions, the LangmuirSchaefer
(LS) lm of 4-CUMC18, which has larger steric hindrance,

Figure 85. (A) Molecular structures of an achiral disk-shaped molecule


having one imidazole unit (118). (B) CD spectra of the supramolecular
gels obtained in dierent batches at 20 C. (C) (a) TEM and (b) AFM
images of twisted nanobers formed from the assembly of 118.
Reprinted with permission from ref 282. Copyright 2010 American
Chemical Society.

twists, as conrmed by a series of CD spectral measurements and


SEM studies. Importantly, the overcrowded stacking of the
cinnamate rings within the assembly plays a very important role
in the spontaneous symmetry breaking and production of an
uneven number of helical nanostructures with dierent handedness.284 Furthermore, this phenomenon may generally occur in
the gel systems of BTA derivatives. It is expected that many of the
other BTA derivatives will show similar properties.
5.4. Air/Water Interface and LB Films

The aggregation of adjacent molecules plays an important role


during the self-assembly of achiral molecules with the emergence
of the supramolecular chirality. The initial stacking at a certain
angle from the neighboring molecules can produce a chiral bias,
and this dislocation can be left-handed or right-handed. If this
chiral bias can be further grown or amplied, chiral structures
with left- and right-handedness are obtained. Moreover, most
importantly, if the growth rate of the two biases is dierent during
the self-assembly, optical activity of the system can be expected,
with an unequal amount of left-handed or right-handed chiral
nanostructures. This type of symmetry breaking is sometimes
observed in the aforementioned processes of self-assembly in
solution and gel systems. However, when the assembly of achiral
molecules was studied at the air/water interface, it was found that
this kind of symmetry breaking is quite general in nature. Thus,
achiral molecules could be fabricated into optically active

Figure 86. Formation of optically active supramolecular gels with chiral nanostructures and optical activity by the hierarchical self-assembly of achiral
119. Reprinted with permission from ref 284. Copyright 2014 John Wiley & Sons.
7354

DOI: 10.1021/cr500671p
Chem. Rev. 2015, 115, 73047397

Chemical Reviews

Review

Figure 87. (A) Chemical structures of NpImC17 (120), and the model for the formation of the chiral NpImC17Ag(I) coordination assemblies. (B)
Circular dichroism (CD) (A) and UV (B) spectra of a 10-layer NpImC17 LB lm transferred from the pure water surface (a) and a 20-layer Ag(I)
NpImC17 LB lm (b). Reprinted with permission from ref 39. Copyright 2003 American Chemical Society.

Figure 88. (A) Structure of achiral amphiphilic barbituric acid 121. (B) AFM images of one-layer LB lms deposited at various surface pressures at 20 C
(a) and (b) 7, (c) 20, and (d) 30 mN/m after the inection point. Reprinted with permission from ref 285. Copyright 2004 American Chemical Society.

showed supramolecular chirality, while 7-CUMC18 did not


(Figure 89).287
The azobenzene derivatives can easily undergo transcis
isomerization; thus, the aggregation behavior of two azobenzene
isomers is often dierent. Liu et al. found that the trans form of
the achiral azobenzene derivative 4-octyl-4-(5-carboxypentamethyleneoxy) azobenzene (trans-C8AzoC5, 124) showed a strong
Cotton eect in its LB lms, while the assemblies of cis-C8AzoC5
did not (Figure 90).288
The aggregation of porphyrin derivatives can be one of the
most attractive models for understanding supramolecular
assemblies. Dierent molecular packing of porphyrins can lead
to H or J aggregation, which have dierent properties. Achiral
porphyrins can also assemble into helical nanostructures with
supramolecular chirality. For example, through the air/water
interfacial assembly of TPPS (67) with achiral positively charged
amphiphiles, Zhang and Liu et al. obtained chiral assemblies, as
shown in Figure 91. The emergence of the supramolecular
chirality could be due to the helical stacking of the TPPS units in
a conned two-dimensional interface.289
The packing mode of the supramolecular assembly of achiral
porphyrins can be modulated upon protonation of their central
nitrogen atoms. If such protonation was performed in situ at the
air/water interface, many achiral porphyrins can assemble into

Figure 89. (A) Molecular structures of achiral coumarin derivatives. (B)


CD spectra of 4-CUMC18 (a, c) and 7-CUMC18 (b) LS lms
transferred from the water surface at 20 mN/m. Reprinted with
permission from ref 287. Copyright 2007 American Chemical Society.

7355

DOI: 10.1021/cr500671p
Chem. Rev. 2015, 115, 73047397

Chemical Reviews

Review

Figure 90. (A) Structure of 4-octyl-4-(5-carboxypentamethyleneoxy)azobenzene (trans-C8AzoC5, 124). (B) Formation of optically active
and inactive supramolecular assemblies from trans-C8AzoC5 (top, 124)
and cis-C8AzoC5 (bottom), respectively. Reprinted with permission
from ref 288. Copyright 2006 American Chemical Society.

Figure 91. (A) CD spectra of 20-layer LS lms of (a) ODA, (b) CTAB,
and (c) DOAB with TPPS transferred at 30 mN/m. (B) Schematic
illustration of TPPS/amphiphile coassemblies and the J aggregation of
TPPS. Reprinted with permission from ref 289. Copyright 2003
American Chemical Society.
Figure 93. (A) Structure of amphiphilic diacetylene derivative (tricosa10,12-diynoic acid, TDA, 125). (B) UVvis and CD spectra of LB lms
deposited from in situ photopolymerized PDA lms (a) from pure water
and (b) from Cu(NO3)2. (C) TEM images of PDA lms in situ
polymerized on pure water. Reprinted with permission from ref 292.
Copyright 2002 The Royal Society of Chemistry.

optically active LB lms, as illustrated in Figure 92.290 Similarly to


these protonated porphyrins, introduction of the metal ions into
the centers of the porphyrin can also cause some of the achiral
porphyrins to form chiral assemblies at the air/water interface.291
Generally, the chiral assemblies obtained from achiral
molecules at the air/water interface are not so stable. Thus, if
the air/water interface is used as the platform to form the chiral
assemblies rst and then introduce more covalent bonds to the
systems, the stability can be increased. One of the best ways is the
transform through the topochemical polymerization. For
example, Liu at al. found that achiral amphiphilic diacetylene
derivatives 125 can form chiral assemblies at the air/water
interface, and these diacetylene derivatives can be polymerized
upon photoirradiation. Interestingly, the photopolymerized
organized molecular lms of polydiacetylene showed strong
optical activity as well as helical nanostructures (Figure 93).292
Zou et al. synthesized dierent achiral azobenzene-substituted
diacetylene monomers (Figure 94) and studied the air/water
interfacial assembly of these molecules.293 The results show that
only the achiral diacetylene monomer containing one hydrophobic chain can form chiral assemblies at the air/water interface.
When these LB lms of compounds 126 and 127 were irradiated

Figure 94. Molecular structures of the three azobenzene-substituted


diacetylene monomers.

Figure 92. Protonation of achiral water-insoluble free-base porphyrins can lead to optically active supramolecular assemblies at the air/water interface.
Reprinted with permission from ref 290. Copyright 2007 John Wiley & Sons.
7356

DOI: 10.1021/cr500671p
Chem. Rev. 2015, 115, 73047397

Chemical Reviews

Review

Figure 95. (A) Molecular structures of dierent porphyrin derivatives. (B) Possible stacking of the porphyrin TPPA3 (133) and TPPA0 (129) assembly
on the air/water interface. Reprinted with permission from ref 296. Copyright 2008 The Royal Society of Chemistry.

and Guo et al. designed a series of TPP-based achiral porphyrins


with dierent alkyl chains and hydrophilic substituents and
investigated the supramolecular chirality within the LB lms as
well as in situ in the monolayers. Depending on the number of
alkyl chains, dierent supramolecular chirality can be detected
from the LB lms of these porphyrin derivatives (Figure 95).296
Furthermore, the production of the supramolecular chirality in
the monolayers of these porphyrins at the air/water interface was
studied by means of second-harmonic generation linear
dichroism (SHG-LD). Interestingly, achiral TPPA2a (131) was
demonstrated to form chiral assemblies by in situ SHG-LD
measurements even though the LS lm did not have clear CD
signals. This is because the SHG-LD can provide more sensitive
signals than conventional CD spectra. In addition, a subtle
interaction at the air/water interface can aect the supramolecular chirality in the monolayers, which can be monitored
by the SHG-LD. For instance, using the SHG-LD technique, it
was found that Cu2+ ions in the subphase enhance the
supramolecular chirality of the TPPA2a monolayer, while Zn2+
ions inhibit the formation of chiral assemblies (Figure 96).297

by left- or right-handed circularly polarized UV light (CPUL),


supramolecular chirality with polymerization of diacetylene
groups can be detected. Interestingly, for the LB lms formed
by compound 128 containing two hydrophobic diacetylene
chains, polymerization cannot be achieved, even though
supramolecular chirality can be detected from the CD signals
of azobenzene chromophores.
In addition to amphiphilic diacetylenes, the chiral polymer
from achiral phthalocyanine derivatives was also achieved. The
achiral phthalocyanine building block silicon
2,3,9,10,16,17,23,24-octakis(octyloxy)-29H,31H-phthalocyanine dihydroxide (Pc 1) was spreading at the air/water interface
to form a monolayer and subsequently transferred onto solid
substrates to produce LS lms with optical activity. Interestingly,
upon heating at 180 C in a high vacuum for 10 h, the chiral
assemblies within the LS lms could be converted into chiral
covalent polymers with the CD signals increased signicantly.294
This result indicated that the transform from the noncovalent to
the covalent bond not only increased the stability of the chiral
assemblies but also amplied the supramolecular chirality of the
systems in these interfacial lms.
In these above examples, for the supramolecular chirality
formed by air/water interfacial assembly of achiral molecules, the
CD spectra have always been measured in the corresponding LB
lms. Thus, one may doubt whether the supramolecular chirality
originated from the deposition process for making LB lms or
merely originates from the assembly of achiral molecules on the
water surface. Therefore, when achiral molecules were placed on
the water surface, the in situ measurement of the supramolecular
chirality could be very important. Recently, the SHG-LD
technique has been developed for detecting supramolecular
chiral assemblies in situ on water surfaces.295 It was conrmed
that the supramolecular chirality is produced from the assembly
of achiral molecules at the air/water interface. In addition, some
subtle eects of the chirality change of the systems can also be
detected from in situ measurements. For example, Liu, Wang,

Figure 96. In situ supramolecular chirality from the assembly of achiral


TPPA2a (131), which can be modulated by adding dierent metal ions
to the subphase. Reprinted with permission from ref 297. Copyright
2014 American Chemical Society.
7357

DOI: 10.1021/cr500671p
Chem. Rev. 2015, 115, 73047397

Chemical Reviews

Review

5.5. Controlling Handedness of Supramolecular Chirality

sulfophenyl)porphyrin (135). Within the assemblies, the achiral


porphyrins were found to undergo J aggregation. In addition,
AFM measurements on the stirred solutions showed helical
nanoribbons, which indicated supramolecular chirality at the
nanoscale (Figure 98).299
With regard to vortex motion or stirring introduced into the
supramolecular chiral assemblies, achiral porphyrins are
important model molecules. In this context, the interactions
between the aromatic rings play very important roles for forming
chiral assemblies. Aida and co-workers studied dendritic zinc
porphyrins with two carboxylic acid groups (136). These
porphyrins could form J aggregates in CHCl3 through
interactions from the porphyrin rings and also from the aromatic
rings within their dendritic substituents. Moreover, hydrogen
bonding between the carboxylic acid groups can also be a very
important driving force for forming J aggregates. The spin
coating of the J-aggregate solutions could produce very stable
optically active lms. Interestingly, the handedness of the
supramolecular chirality of the lms can be controlled by
changing the direction of the spin coating (Figure 99).248
Aside from spin coating, the stirring of benzene solutions of
achiral dendritic zinc porphyrins with two carboxylic acid groups
(136) can also produce supramolecular chirality. Strong CD
signals can be detected from the porphyrin solution upon
stirring. Interestingly, when the direction of stirring was changed
from clockwise (CW) to counterclockwise (CCW), the opposite
CD signals were detected. For this system, the nanobers formed
from the assembly of porphyrins via J aggregation play very
important roles. Some of the observed chiroptical activity could
also originate from the macroscopic helical alignment of
nanobers (Figure 100).300
For achiral porphyrins in aqueous solution under stirring,
Purrello et al. carefully studied the changes of CD spectra of
TPPS (67) aqueous solutions in cuvettes upon stirring.301 These
authors found that by stirring in dierent directions the
protonated TPPS could form two dierent J aggregates. The
results showed that the CD signal of the solution inverted with an
increasing in stirring intensity when the direction of stirring was
altered (Figure 101B). Remarkably, with stirring, assemblies of
the porphyrin were deposited on the cuvette wall, and the
chirality of the assemblies on the solid surface were dierent than
those in solution. However, the directions of the CD signals of
the porphyrin aggregations on the cuvette wall were also found to
be dependent on the stirring direction.301
In some cases, strong CD signals can be detected from
assemblies of achiral molecules, but the supramolecular chirality
of these assemblies is still in doubt. Linear polarization
properties, such as linear dichroism (LD) and linear
birefringence (LB), can hamper the actual circular dichroism
(CD). Images of the true CD signal can be obtained using
Mueller matrix spectroscopy (MMS). For some assemblies,
formed by using achiral molecular building blocks, using MMS
may overcome the problems related to linear polarization
properties. Okano and Yamashita et al. developed a twocomponent system to investigate the supramolecular chirality
that originates from the assembly of achiral molecules.303 In this
study, the achiral ionic oligomer (poly[pyridinium-1,4-diylimino-carbonyl-1,4-phenylene-methylene chloride], 138) formed
hydrogels at high concentration. At lower concentrations in
water, the achiral ionic oligomer could form brous nanostructures with supramolecular chirality with vortex stirring. To
remove the inuence of the linear polarization properties for
identifying the supramolecular chirality, another achiral dye

Constructing supramolecular chiral assemblies from achiral


molecular building blocks is a fascinating possibility. A further
step would be to control the handedness of supramolecular
chirality containing only achiral molecular building blocks.
Unfortunately, there are still very few methods available for
controlling handedness. However, we can gain hints from several
important cases. Herein, we focus on the eects of mechanical
force and circularly polarized light.
5.5.1. Vortices and Spin Coating. For the formation of
supramolecular chiral assemblies from achiral molecular building
blocks, the macroscopic force eld may play a very important
role. It is worth mentioning that the handedness of chiral
assemblies formed by achiral molecules can be controlled by the
direction of vortex stirring. This is the advantage of symmetry
breaking with vortex stirring or spin coating. Considering that the
chiral information is exhibited below the nanoscale, while
mechanical perturbations occur on macroscopic scales, the
modulation of supramolecular chirality is remarkable.298
With regard to this issue, the rst signicant work was
published by Ribo and co-workers.249 The molecular building
blocks for these studies were various achiral diprotonated mesosulfonatophenyl-substituted porphyrins. A vortex motion with
either clockwise or anticlockwise direction was generated using a
rotary evaporator with dierent directions of rotation. Upon
rotary evaporation, the achiral diprotonated meso-sulfonatophenyl-substituted porphyrins could form assemblies with
optical activity. Interestingly, the chirality sign of these
assemblies can be selected by vortex motion during the
aggregation process, as conrmed by circular dichroism (CD)
spectra. In this context, the aggregation of the porphyrins could
be modulated to show dierent supramolecular chirality by
changing the rotational direction of the rotary evaporator (Figure
97).
The supramolecular chirality from the assembly of achiral
meso-sulfonatophenyl-substituted porphyrins under vortex motion or stirring can be detected from the CD spectra and
observed directly from TEM and AFM measurements. Ribo et al.
studied stirred solutions of 5-phenyl-10,15,20-tris(4-

Figure 97. Achiral diprotonated meso-sulfonatophenyl-substituted


porphyrins form optically active assemblies upon rotary evaporation.
Reprinted with permission from ref 249. Copyright 2001 The American
Association for the Advancement of Science.
7358

DOI: 10.1021/cr500671p
Chem. Rev. 2015, 115, 73047397

Chemical Reviews

Review

Figure 98. (A) Structure of 5-phenyl-10,15,20-tris(4-sulfophenyl)porphyrin (135). (B) AFM images of assemblies of 135 showing the onset of folding
in stagnant and stirred solutions. Reprinted with permission from ref 299. Copyright 2006 John Wiley & Sons.

Figure 99. (A) J aggregation of achiral dendritic zinc porphyrin (136) and macroscopic chirality from spinning. (B) Mechanism for the formation of
chiral assemblies. Reprinted with permission from ref 248. Copyright 2004 John Wiley & Sons.

molecule, Congo Red (CR, 137), was added to the system. As a


result, the real supramolecular chirality of the assemblies of
achiral ionic oligomers, i.e., the chiral information, was
transferred to the Congo Red. In this context, a strong CD
signal from the Congo Red can be detected by Mueller matrix
spectroscopy. Amazingly, due to the monosignate shape of the
UVvis spectra of Congo Red, which suggested the absence of
chromophore coupling, the chirality of Congo Red was
attributed to its single chromophore, instead of its aggregates.302
These results suggest that the supramolecular chirality can in fact
be generated from an assembly of achiral molecules (Figure 102).
Similar chiral assemblies from achiral two-component
molecular building blocks were studied using circularly polarized
luminescence. For this study, the concentration of the achiral
ionic oligomer (138) was increased to form hydrogels, and
Rhodamine B dye (139) was embedded into the hydrogel.
Vortex stirring was performed during the slow cooling from the
sol to the gel. The supramolecular chirality was studied by
circularly polarized luminescence (CPL) from Rhodamine B dye,
and supramolecular chirality was detected in these hydrogels.
When the samples were heated to change the gel to a solution,
the resulting supramolecular chirality disappeared. Interestingly,
these results also show that the sense of the CPL could be
controlled by switching the stirring direction from clockwise
(CW) to counterclockwise (CCW) (Figure 103).303
The mechanism of the formation of chiral assemblies formed
by achiral molecular building blocks with vortex stirring has been

Figure 100. Molecular structure of achiral dendritic zinc porphyrin


(136) and the corresponding supramolecular chiral assembly generated
from vortex ows. Reprinted with permission from ref 300. Copyright
2007 John Wiley & Sons.
7359

DOI: 10.1021/cr500671p
Chem. Rev. 2015, 115, 73047397

Chemical Reviews

Review

Figure 103. (A) Structure of achiral ionic oligomer (138) and


Rhodamine B dye (139). (B) (a) Photoluminescence (PL) spectra
(ex = 520 nm) of Rhodamine B (139) (1.6 105 M) in an aqueous
solution of 138 (0.6 wt %). (B) Circularly polarized luminescence
spectra (CPL) of the solution with clockwise (CW, red) and
counterclockwise (CCW, blue) stirring at 1000 rpm and unstirred
(black). Reprinted with permission from ref 303. Copyright 2011 John
Wiley & Sons.

Figure 101. (A) Schematic structure of the protonation of TPPS. (B)


Schematic representation of the possible eects of the stirring on a
racemate (A). CW (B) and CCW (C) stirring favor and J
aggregates, respectively. Reprinted with permission from ref 301.
Copyright 2010 John Wiley & Sons.

porphyrins (Figure 104A) can be a hierarchical noncovalent


polymerization process preceded by a critical nucleation stage. In
this case, the primary nucleation is a very slow process, while the
secondary nucleation stage is much faster. During the primary
nucleation process, signicant enantiomeric excesses can be
obtained from the formation of a few primary nuclei, and the
vortex stirring controlled the secondary nucleation by assisting

Figure 102. Chiral induction to an achiral molecule (Congo Red dye,


137) from the chiral assembly induced by an achiral oligomer (138) was
demonstrated by Mueller matrix spectroscopy analysis of the optical
polarization properties. The true CD image of the clockwise (top) and
counterclockwise (bottom) stirred solution of achiral ionic oligomer
138 are shown. Reprinted with permission from ref 302. Copyright 2011
John Wiley & Sons.

Figure 104. (A) Molecular structures of dierent achiral diprotonated


meso-sulfonatophenyl-substituted porphyrins. (B) CD spectra of
H4TPPF5S3 (140) (10 m, HCl 0.1 M) J aggregates obtained under
vigorous-shaking conditions (6 h, full line) and in magnetically stirred
conditions (24 h, dashed line). Reprinted with permission from ref 304.
Copyright 2012 John Wiley & Sons.

studied by Ribo et al. Their results show that the assembly of


dierent achiral diprotonated meso-sulfonatophenyl-substituted
7360

DOI: 10.1021/cr500671p
Chem. Rev. 2015, 115, 73047397

Chemical Reviews

Review

the growth of the primary chiral nuclei.304 This mechanism can


be demonstrated from the dierence between the CD spectra
measured under vigorous shaking and magnetic stirring (Figure
104B). Thereby, chiral assemblies can be obtained from achiral
building blocks.
Although vortex stirring can help achiral molecules to form
chiral assemblies, both rotational and magnetic forces provided
more ecient modulation of the supramolecular chirality during
the assembly of achiral molecules. This issue is fully
demonstrated by Scolaro et al. using the aggregation of an
achiral porphyrin, tris(4-sulfonatophenyl) phenylporphyrin
(TPPS3, 135), to form chiral assemblies under both rotational
and magnetic forces.305 Interestingly, the inuence of the
magnetic forces can be tuned to an eective gravity which
plays a very important role, while the magnetic orientation of the
aggregates is also essential (Figure 105). Moreover, this study
also suggested that application of rotational and magnetic forces
during the primary nucleation process is sucient to form chiral
assemblies.

Figure 106. (ad) Self-assembly leading to the formation of a twodimensional emulsion featuring chirally resolved domains. transAzobenzene surfactant molecules (b, 141), resulting from the
spontaneous isomerization of the cis isomer (a), self-assemble at the
air/water interface into segregated chirally resolved circular domains
surrounded by the cis-rich matrix (d); (e) molecular structures of achiral
trans-azobenzene surfactant 141 and its chiral analogues 142 and 143.
(f) Coupling between the chiral modier and the vortical ow.
Reprinted with permission from ref 306. Copyright 2012 Nature
Publishing Group.

molecular assemblies containing only achiral molecular building


blocks were irradiated by circularly polarized light, chirality can
be introduced into the systems.
Oriol et al. studied liquid crystals with chiral organization
formed by achiral molecules upon irradiation with circularly
polarized light (Figure 107). The achiral building block used was
a polymer with methoxyazobenzene groups in the side chain.
When thin lms of the nematic glassy phase of this polymer were
irradiated with 488 nm circularly polarized light (CPL), a chiral
arrangement of the azobenzene groups with helical nanostructures resulted from the selective reection of visible light.
Furthermore, the CPL-induced supramolecular chirality in the
polymer was conrmed by CD spectra and vibrational circular
dichroism (VCD) spectra.201
Zou et al. synthesized an achiral amphiphilic azobenzene
derivative (144), which can assemble into a monolayer on the
surface of water (Figure 108A). Supramolecular chirality with
helical packing of azobenzene derivatives can be introduced into
these monolayers by irradiation with CPL. The resulting
handedness of the supramolecular chirality of these monolayers
was controlled by the handedness of the CPL. This is an excellent
example that supramolecular chirality from an achiral component
be controlled by CPL. Furthermore, these chiral azobenzene
monolayers can be used as a chiral template, which supports the
contention that the diacetylene derivatives can form chiral
polydiacetylene under normal UV irradiation (Figure 108B).309
Moreover, Zou et al. also prepared discotic hydrogen-bonding
complexes of diacetylene (DA) units. When these complexes
were irradiated by circularly polarized ultraviolet light within the
liquid-crystal phase, helical polydiacetylene can be obtained
(Figure 109A).310 Most interestingly, by using linearly polarized
light irradiation together with a magnetic eld, the enantioselective polymerization of these complexes can also be achieved
within the liquid-crystal phase (Figure 109B).311 In this case, the

Figure 105. Chiral self-assembly of TPPS3 under rotational and


magnetic forces, showing the relationship between the observed chirality
and the applied physical forces. Reprinted with permission from ref 305.
Copyright 2012 Nature Publishing Group.

Nevertheless, supramolecular chirality originating from


assemblies of achiral molecular building blocks upon vortex
stirring has been generally acknowledged. On the other hand, it is
well known that some chiral dopants can coassemble with achiral
molecular building blocks to form chiral supramolecular
aggregations. When the chiral species and the vortex stirring
meet together, which eect will be dominate? Sagues and coworkers studied the chiral competition between vortex rotating
and chiral dopants.306 In their study, achiral trans-azobenzene
surfactant 141 was used as the primary building block, while its
chiral analogues 142 and 143 were the chiral dopants. Vortex
rotating was performed to form monolayer assemblies, and the
supramolecular chirality was studied using Brewster angle
microscope (BAM) measurements. The results showed that
the inuence of the vortex rotation can be comparable to that of
the chemical induction processes. However, the inuence of the
vortex rotation can be more easily controlled by changing the
direction or speed of stirring (Figure 106).
5.5.2. Circularly-Polarized Light. Circularly polarized light
(CPL) can be regarded as a type of energy that contains chiral
information. When achiral supramolecular assemblies are
irradiated by circularly polarized light, chiral information from
the illumination can be transferred to the systems to form
supramolecular chiral assemblies.
For more than a century, scientists have considered CPL as a
possible cause of natural homochirality. Because CPL has been
observed in star formation, CPL could be the source of chiral
information during the origins of life.307,308 When supra7361

DOI: 10.1021/cr500671p
Chem. Rev. 2015, 115, 73047397

Chemical Reviews

Review

control on the supramolecular chirality in self-assembled


systems.
5.5.3. Surface Pressure. Indeed, many supramolecular
chiral assemblies can be obtained from dierent achiral molecular
building blocks via air/water interfacial assembly. However, for
this method, a great remaining challenge is the control of the
macroscopic chirality of these systems. Normally, the handedness of the chirality of the assemblies obtained from the air/water
interfacial assembly from achiral molecular building blocks is
random. In order to control the macroscopic chirality of air/
water interfacial assembly from achiral molecules, a new method
using unilateral compression geometry has been developed. In
this approach, only one of the two movable Langmuir barriers
was used for the compression. Using this kind of unidirectional
compression, Chen and Liu et al. found that the assemblies
deposited from the mirror regions of the LB trough can display
mirror-image macroscopic chirality.312 During the compression,
vortex-like ows are suggested to be generated, and the direction
of this compression-generated vortex-like ow can determine the
macroscopic chirality of the formed assemblies (Figure 110).
5.6. Self-Assembly of Racemic Systems

Molecular or supramolecular chirality plays a very important role


in many self-assembly processes. In contrast to constructing
chiral assemblies from pure achiral building blocks, the selfassembly manner of racemic systems is unique and largely
dependent on the mixing of chiral enantiomers with dierent
handedness.
First, the mixing of an enantiomer and its mirror image may
meet with problems such as phase separation, so that the
situation of these systems can be very complicated. In general, for
self-assembly of racemic systems, regardless of the chiral
molecules or the chiral nanostructures, some enantiomers can
preferentially aggregate with themselves, while other enantiomers might prefer to form complexes with their mirror images.
Therefore, self-recognition can be established when some
enantiomers recognize themselves and form homochiral
assemblies, while in the case of an R enantiomer forming a
complex with its S enantiomer, this kind of self-sorting has been
regarded as self-discrimination to form heterochiral coassemblies.313317
For the self-assembly of racemic systems, in most cases, there
are much more possibilities for forming homochiral assemblies
than that of forming heterochiral assemblies. However, because
the energy dierence between the enantiomers with dierent
handedness is quite small, phase separation in racemic systems is
not very signicant.72
Wurthner et al. studied the chiral self-sorting of perylene
bisimide (PBIs) assemblies.318 In this case, dierent chiral PBI
molecules containing oligoethylene glycol bridges have been
synthesized, and the coassembly of these enantiomers with
dierent handedness was investigated. The results showed that
chiral self-recognition always prevails over self-discrimination for
these PBIs assemblies. The coassembly of this system tends to
form homochiral aggregates (Figure 111).
Percec et al. synthesized chiral hat-shaped dendronized
cyclotriveratrylene (CTV) and studied the racemic assembly of
these molecules. The results showed that these molecules can
form heterochiral assemblies at low temperature. When these
systems were heated below 60 C for 2 h, homochiral assemblies
formed upon self-sorting (Figure 112).319
Although chiral self-sorting can be very signicant for a
racemic assembly, systems containing enantiomers with dierent

Figure 107. Model for circularly polarized-light introduced chirality in


azomaterials. Reprinted with permission from ref 201. Copyright 2007
John Wiley & Sons.

Figure 108. (A) Molecular structures of achiral amphiphilic azobenzene


derivative 144 and achiral diacetylene derivative 125. (B) Schematic
illustrations of the generation of chirality from PTDA/DBA hybrid lms
with the azobenzene-containing monolayer irradiated by left- and righthanded CPL. Reprinted with permission from ref 309. Copyright 2011
The Royal Society of Chemistry.

handedness of the helical polydiacetylene chains can be


controlled by changing the orientation of the linearly polarized
light and the magnetic eld. The enantioselective recognition of
D- or L-lysine using these helical polydiacetylene assemblies was
achieved. Combination of these physical vectors showed eective
7362

DOI: 10.1021/cr500671p
Chem. Rev. 2015, 115, 73047397

Chemical Reviews

Review

Figure 109. (A) Structure of discotic hydrogen-bonding complex of diacetylene (DA) units. (B) Asymmetric polymerization of the corresponding
complex by using linearly polarized light irradiation and a magnetic eld. Reprinted with permission from ref 310. Copyright 2014 The Royal Society of
Chemistry. Reprinted with permission from ref 311. Copyright 2014 Nature Publishing Group.

Figure 110. Schematic illustrations of the chirality selection


phenomenon induced by compression-generated vortex-like ow.
Reprinted with permission from ref 312. Copyright 2011 John Wiley
& Sons.

Figure 111. (A) Structures of chiral perylene bisimides (PBIs)


containing oligoethylene glycol bridges (146). (B) Energy diagram of
the dierent supramolecular diastereomers of PBIs. Reprinted with
permission from ref 318. Copyright 2011 American Chemical Society.

handedness can still show many new features, of which the


nanostructures and properties are dierent from corresponding
pure chiral assemblies.67 For example, for the formation of
supramolecular gels, the racemates are usually believed to be
poor gelators. However, in some special cases, only racemates
were found to form supramolecular gels while the corresponding
enantiomer cannot.320
For example, Yamaguchi et al. found that homochiral
pseudoenantiomeric ethynylhelicene oligomers (147) cannot
form organogels, while the mixture of P oligomer and M
oligomer can form two-component organogels in toluene.

Changing the ratio of the pseudoenantiomers produced various


gel systems (Figure 113).321
With the exception of the formation of supramolecular gels,
the self-assembly of racemic molecular building blocks can also
improve the macroscopic properties of soft matter. For example,
the mechanical properties of peptide hydrogels can be improved
upon racemic self-assembly. Schneider et al. found that the
hydrogels prepared from racemic -hairpins showed nonadditive, synergistic enhancement in material rigidity compared
to gels prepared from either pure enantiomer. Therefore, racemic
7363

DOI: 10.1021/cr500671p
Chem. Rev. 2015, 115, 73047397

Chemical Reviews

Review

Figure 112. Chiral self-sorting during supramolecular helical organization of hat-shaped molecules. Reprinted with permission from ref 319. Copyright
2014 American Chemical Society.

Figure 114. (A) (a) Assembly mechanisms for enantiomeric peptides


leading to the formation of a brillar network that denes hydrogelation.
(b) Sequences of enantiomers MAX1, DMAX1, and the nonisomeric
control peptide. D-Amino acid residues are italicized. (B) Dynamic time
sweep rheological data measuring the storage moduli of 1 wt %
hydrogels containing pure MAX1 (), 3.1 MAX1.DMAX1 (), 1.1
MAX1.DMAX1 (), 1.3 MAX1.DMAX1 (), and pure DMAX1 ().
Reprinted with permission from ref 322. Copyright 2011 American
Chemical Society.

Figure 113. (A) Structures of pseudoenantiomeric ethynylhelicene


oligomers. (B) Minimal thermoreversible gelation concentration in
millimolar for 1.1 mixtures of (M)-n and (P)-n (n = 16) in toluene. S =
soluble at room temperature (solubility > 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 dierent 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 dierent
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, 73047397

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 dierent 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 dierent molar ratios further changed the
nanostructures consecutively from helical nanotubes to nanotwists to at nanoplates (Figure 117).53
In most cases, the self-assembly of chiral molecules can lead to
helical nanostructures, while racemates usually assemble into at
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 at nanostructures, even though both LAlaC17 and D-AlaC17 can form gels in dierent 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 NH, 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 dierent materials within diverse scales.11,324326


Herein, we address the most prominent elds 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, 73047397

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 dierent enantiomers
and another chiral molecule can produce totally dierent results.
These dierent interactions can be detected by spectral
measurements or determination of their crystal structures,318,327329 representing the chiral recognition. Thus far,
studies on hostguest chemistry related to chirality have focused
on chiral recognition,330337 and many excellent works have
been published. For example, work on chiral recognition using
cyclodextrins is very famous and has been extensively
reviewed.338346
On the other hand, chiral recognition not only describes the
interactions between dierent 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.347349 Thus, when
molecules with contrary chirality interact with chiral supramolecular assemblies, the chiral recognition shows some new
features, which are dierent from the case of chiral recognition
between dierent chiral molecules. For example, when molecules
with dierent 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 solgel
transformation or color changes of the supramolecular gels, are
visible to the naked eye.350,192
Although chiral recognition related to the hostguest
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 signicant 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 cyclodextrin 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
uorescence of rare earth complexes, the recognition of small
chiral molecules by this supramolecular polymer was achieved
based on both uorescence and circular dichroism spectral
changes.353
One of the most important advantages of the chiral
recognitions from dierent chiral soft matters can be the
alteration of their macroscopic properties, which also can be
7366

DOI: 10.1021/cr500671p
Chem. Rev. 2015, 115, 73047397

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 (nal concentration;
L, TbIII, and D/L-tartaric acid, 10 m, -CD, 300 m). (dotted line) L +
-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 Schi base moieties and long alkyl chains (oSLG, 8a and p-SLG, 8b). The self-assembly and gelation
properties of these amphiphiles with dierent metal ions have
been investigated in many organic solvents. In particular, adding
metal ions to the organogels can signicantly change the selfassembled nanostructures and the spectral characteristics of
these systems. Thus, Cu2+ can transform the nanober gel into a
chiral twist, while adding Mg2+ ions enhanced the uorescence 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 uorescence quenching
processes of these organogels can be totally dierent (Figure
123).158
Liu et al. designed and synthesized a series of amphiphilic
molecules containing both glutamide moieties and long alkyl
chains with dierent 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 dierent 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 uorescent metallogels with optical activity upon
gelation in several organic solvents. The recognition of the small
chiral organic molecules by these uorescent metallogels can be
detected by the changes in their CD spectra or uorescence
spectra. For example, metallogels formed by Zn2+/HQLG (10)
complex showed a totally dierent uorescent 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 dierent amino acids by assemblies thereof (Figure 125A). gTPP/Zn (150) can form organogels in dierent organic solvents.

identied 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 solgel transformation. The work of Pu et al. is the rst 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 solgel 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 solgel 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, 73047397

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 M) with and without L- and D-His-OMe (50 M) in
cyclohexane at 20 C. (C) Fluorescence spectra of the g-TPP/Zn (150)
(50 M) assembly with and without L- and D-His-OMe (50 M) 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 Schi 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 bers, 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 diusion of R amines
and displacement of the original S amines from the solid-like
bers 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 dierent -amino acid derivatives can be
dierentiated 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
dierent CD and uorescence 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 dierent 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, 73047397

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 congurations 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 dierent devices for
chiral sensing.
Qu et al. constructed electrochemical DNA sensors for the
chiral sensing. In this system, DNA molecules modied 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 oers 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-nger-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 dicult
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 ecient 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 uorophores have
been synthesized. This is the result of the salicylaldehyde unit
which can form a complex with amino alcohols with dierent
chirality and subsequently change the conformation of the
molecules. The chiral sensing can be determined from the
changes in the CD and uorescence spectra. By using this system,
the absolute conguration 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.362367
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-decient uorinated porphyrin tweezer
(152) and demonstrated that it is a good chiral sensor for the
discrimination of many dierent 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 conguration of a variety of erythro and threo
7369

DOI: 10.1021/cr500671p
Chem. Rev. 2015, 115, 73047397

Chemical Reviews

Review

Figure 128. (A) Molecular structures of the secondary terephthalamide derivatives containing four aryl blades (153) and their dierent 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, 73047397

Chemical Reviews

Review

Figure 131. (A) Four-component reversible covalent assembly for secondary alcohol binding. OTf is triuoromethanesulfonate (triate). (B)
Exploration of four-component assembly for chirality sensing and ee determination. CD spectra of 1-phenylethanol-induced assembly with dierent ees
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 conguration 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 dierent
ee values (Figure 131B).373
A chiral sensor with the ability to dierentiate many dierent
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,375380 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 dierences at pictogram
quantities. In this study, the optical excitation of plasmonic
planar chiral metamaterials can generate superchiral electromagnetic elds, 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 dierences were
observed for proteins with high -sheet content. The system
7371

DOI: 10.1021/cr500671p
Chem. Rev. 2015, 115, 73047397

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) Inuence of the adsorbed proteins hemoglobin, -lactoglobulin, and thermally
denatured -lactoglobulin on the CD spectra of the PCMs. (C) Hemoglobin (top) and -lactoglobulin (bottom) (-helix, cyan cylinder; -sheet,
ribbons), shown adopting a well-dened arbitrary structure with respect to a surface. The gure illustrates the more anisotropic nature of adsorbed lactoglobulin. 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 eld 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 specic
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.382384 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
hostguest complexes,386 LB lms,387389 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.391396
A supramolecular chiroptical switch based on an amorphous
azobenzene polymer (156) has been constructed by Kim et al.
When the thin lms 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 conrmed 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 rst reversible supramolecular
chirality switch based on chiral polythiophene aggregates. When
copper(II) triuoromethanesulfonate [Cu(OTf)2] was added to
chiral aggregates of chiral regioregular polythiophene in a
chloroformacetonitrile 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 dierent 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 lms 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 exibility
7372

DOI: 10.1021/cr500671p
Chem. Rev. 2015, 115, 73047397

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 lms
via an air/water interfacial assembly. Supramolecular chiral
switches based on the LS lms of achiral TARC18 were
fabricated by alternately exposing the lm 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 lms 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 lms. M and P chiralities of the lms were formed by
chance. Reprinted with permission from ref 387. Copyright 2006 John
Wiley & Sons.

their well-known work on light-driven molecular motors.399401


In this context, the modied 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
dierent 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 lms 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 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 lms 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 lms of azobenzene-substituted
diacetylene (NADA, 158). Although NADA is achiral, the
NADA LB lms show supramolecular chirality. Moreover, the
assemblies within the LB lms can be polymerized with
photoirradiation. When left- and right-handed circularly
polarized ultraviolet light (CPUL) was applied, polymerized
NADA (PNADA) LB lms with dierent chirality were obtained,
as conrmed by the corresponding CD spectra from both
azobenzene chromophores and polydiacetylene (PDA) chains.
Interestingly, these polymerized LB lms NADA (PDA LB lms)
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 eld 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 polymers 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 dierent 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 binaphthalenebipyridinium guests together with cucurbituril hosts have been developed by Venturi and
7373

DOI: 10.1021/cr500671p
Chem. Rev. 2015, 115, 73047397

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 dierent length alkyl chains were designed and
synthesized (159). In aqueous solution, these chiral binaphthalenebipyridinium 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 UVvis and in
the CD spectra upon alternative UVvis 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 dierent
chiral homogeneous catalysts and heterogeneous catalysts has
become a signicant research goal.405411 For example,
coordination complexes containing metal ions and chiral ligands
have been widely investigated as highly ecient chiral catalysts.
Many outstanding reviews have been published examining this
issue.410416 In particular, as a class of very important
coordination complexes with crystalline structures, metal
organic frameworks (MOFs), which have innite 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 modied
to obtain more ecient 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 eld 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 binaphthalenebipyridinium 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 dierent
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, 73047397

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--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 Feringas
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 -dehydroamino acid, enamide, and
itaconic acid derivatives, these reusable heterogeneous asymmetric catalysts can lead to very high reaction rates with excellent
enantioselectivity (9097% 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 ,-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(pentauorophenyl)borane (162). In the DielsAlder
reactions of cyclopentadiene with dierent 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 Feringas 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--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, 73047397

Chemical Reviews

Review

Figure 143. (A) Molecular structure of a ligand containing the 2,2:6,2-terpyridine (tpy) unit and Feringas MonoPhos (160), and schematic
illustration of supramolecular catalysts through orthogonal coordination of two dierent 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 ecient in the
asymmetric catalysis of the unusual DielsAlder reaction
between cyclohexenones and nitrodienes, nitroenynes, or
nitroolens, providing excellent chemo-, regio-, and enantioselectivities (Figure 146).427

Figure 146. Coassembly of chiral amines and poly(alkene glycol)s


showing highly ecient asymmetric catalysis of DielsAlder 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 ,-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 DielsAlder 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, 73047397

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


riboavin units as the main chain (165). Within the polymers,
the 5-ethylriboavinium cations can be reversibly transformed
into 4a-hydroxyriboavins upon hydroxylation/dehydroxylation,
which renders signicant changes in the absorption and circular
dichroism (CD) spectra of the polymers. It is believed that the
face-to-face stacking of the intermolecular riboavinium units
within the polymer produced twisted helical nanostructures with
supramolecular chirality. This optically active polymer containing 5-ethylriboavinium cations was found to eciently catalyze
the asymmetric organocatalytic oxidation of suldes 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 dierent applications. As a polymer, DNA has many
dierent 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 rst 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 DielsAlder 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 purication 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, 73047397

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 aording 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 riboavin units catalyze the asymmetric organocatalytic oxidation of suldes. 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 olens in
an aqueous environment, resulted in good eciency under the
laboratory conditions. For the enantioselective hydration of
enones, the chiral -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
diastereospecic, with the formation of only the syn hydration
product.436
The folding of DNA can produce dierent 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 DielsAlder 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 dierent metal ions. In this case, additional achiral
ligands were not needed for building the catalysts. In an
asymmetric DielsAlder reaction, the complex of human
telomeric G4DNA and Cu2+ ions provided a signicant
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 ,unsaturated 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
FriedelCrafts alkylation with olens 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 FriedelCrafts reaction of ,unsaturated 2-acylimidazoles with heteroaromatic 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 eciency of both
double-stranded DNA and single-stranded DNA with dierent
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
FriedelCrafts 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 151155). For the
asymmetric intramolecular cyclopropanation of -diazo--keto
7378

DOI: 10.1021/cr500671p
Chem. Rev. 2015, 115, 73047397

Chemical Reviews

Review

Figure 151. (A) Asymmetric DielsAlder 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) Cudmbpy/st-DNA-catalyzed FriedelCrafts 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 ,-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 -diazo--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 DielsAlder reactions with G-quadruplex-DNA-based catalysts; the absolute conguration 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
conguration 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 conguration 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 dierent 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, 73047397

Chemical Reviews

Review

Figure 153. Tuning the absolute conguration 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 exible 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 articial 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 DielsAlder 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 modications. These
polypeptides are cyclic peptides formed by intramolecular
disulde 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 exibility in the amino acids of the
polypeptides. In catalyzing DielsAlder and FriedelCrafts
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 dierent supramolecular chirality (Figure


153). In the case of FriedelCrafts reactions and Michael
additions using many dierent 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.
Andreasson et al. studied the asymmetric closing reaction of
dithienylethene derivatives by complexion with DNA upon
photoirradiation. In this study, uorinated 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 signicant 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 articial 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 disulde bridge. (b) DA reaction catalyzed
by cyclic peptide ligand and Cu2+. Reprinted with permission from ref
442. Copyright 2014 John Wiley & Sons.

Figure 155. Enantioselective articial metalloenzymes based on a bovine pancreatic polypeptide scaold, showing catalytic DielsAlder 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, 73047397

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.443448 Chiral
optics and electronics are directly dependent on many dierent
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 microbers with clear screws and favorable
monodispersity from chiral polyaniline (PANI). When these
superhelical microbers were treated with chiral aminohexane
vapor with dierent handedness, very dierent electrical
conductivity in these microbers was detected (Figure
156).449,450

Figure 157. Superhelical conducting microbers 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 rst important work concerning chiral optics was
published by Verbiest et al. in 1998. The authors prepared LB
lms of helicene with dierent supramolecular chirality and
studied the second-order nonlinear optical (NLO) properties of
these LB lms. 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 materials chirality. For example, Fourmigue et al. studied the
electronic conductivity of chiral salts of tetrathiafulvalene
methyloxazoline 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 dierent 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.
Mullen et al. studied eld-eect 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 dierent
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 eld
eect transistors from the assembly of helicene (Figure 160A),
and a highly specic 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, 73047397

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


thiophenephenylene copolymers and found that these
polymers, with dierent molecular structures and aggregation
states, could exhibit red, green, and blue uorescent. Remarkably,
mixing these dierent uorescent 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 dierent 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 dierent 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 eorts are
based on the -conjugated polymers with chiral side chains or
helical aggregated nanostructures that have been reported to
show intense CPL signals.460465
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, dierent supramolecular assemblies
were found to produce opposite CPL spectra.463

Figure 161. Mixture of red, green, and blue uorescent 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
systems.466469
Maeda et al. introduced the BINOLboron moiety to
dipyrrolyldiketones and fabricated the chiral conformation of
-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, 73047397

Chemical Reviews

Review

1,4-benzenedicarboxamide phenylalanine derivative (170) as


supramolecular gelators to construct dierent supramolecular
hydrogels, and the cell adhesion within these supramolecular
hydrogels was studied. It was found that cell adhesion and
proliferation can be inuenced by the chirality of the nanobers.
Thus, the left-handed helical nanobers increased cell adhesion
and proliferation, while the right-handed nanobers decrease cell
adhesion and proliferation. The stereospecic interaction
between chiral nanobers and bronectin plays a critical role in
these eects (Figure 163).476

chromophoric units played a crucial role in the eective


enhancement of chiroptical properties in these self-assembled
structures.470

Figure 163. Schematic representation of the culture of cells in


supramolecular hydrogels and the dierent cell-adhesion and cellproliferation behavior in the enantiomeric nanobrous hydrogels (d:
right-handed helical nanobers; l: left-handed helical nanobers). 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 eects on cell adhesion.
Supramolecular hydrogels formed from the self-assembly of
peptide derivatives or nucleic acid derivatives have been studied
for dierent biological applications.471473 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 eects of amino acid chirality on tripeptide selfassembly and hydrogelation at physiological pH and cytocompatibility in broblast cell culture. In this study, dierent
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 broblasts
in vitro.475
Interestingly, the cell adhesion in the supramolecular hydrogels was found to be dependent on the handedness of the selfassembled nanobers. Feng et al. used the two enantiomers of a

As mentioned above, the handedness of self-assembled


nanostructures can inuence cell proliferation. However, Zouani
et al. demonstrated that helical nanostructures with the same
handedness, but dierent shapes and periodicities show totally
dierent 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 dierent shapes
and periodicities. Interestingly, helical silica nanoribbons with a
specic periodicity of 63 nm (5 nm) helped the specic cell
adhesion and stem cell dierentiation, while silica twists with a
specic 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
dierence could be detected (Figure 163).478
7383

DOI: 10.1021/cr500671p
Chem. Rev. 2015, 115, 73047397

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 dicult 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 dicult. 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 dierent molecules into a complex chiral system in a
cooperative or syndetic way as is accomplished in a living cell.
Nanostructured chiral materials oer many opportunities to
develop entirely new functional materials, which justies 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 eects are fundamental
to biological systems, such as dierent enantiomers that can be a
useful drug and or poisonous depending on their chirality. How
to construct the chiral/biointerface? Therefore, further eorts 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 dierentiation 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 dierent 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. Dierent 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: liumh@iccas.ac.cn.


Notes

The authors declare no competing nancial 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, 73047397

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 lms, supramolecular gels, and soft nanomaterials.

(1) Hegstrom, R. A.; Kondepudi, D. K. The Handedness of the


Universe. Sci. Am. 1990, 262, 108115.
(2) Bai, C.; Liu, M. From Chemistry to Nanoscience: Not Just a Matter
of Size. Angew. Chem., Int. Ed. 2013, 52, 26782683.
(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., Vogtle, 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, 47694774.
(6) Alberts, B.; Bray, D.; Lewis, J.; Ra, 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, 17891816.
(8) Perez-Garca, L.; Amabilino, D. B. Spontaneous Resolution under
Supramolecular Control. Chem. Soc. Rev. 2002, 31, 342356.
(9) Mateos-Timoneda, M. A.; Crego-Calama, M.; Reinhoudt, D. N.
Supramolecular Chirality of Self-Assembled Systems in Solution. Chem.
Soc. Rev. 2004, 33, 363372.
(10) Perez-Garca, 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, 941967.
(11) Hembury, G. A.; Borovkov, V. V.; Inoue, Y. Chirality-Sensing
Supramolecular Systems. Chem. Rev. 2008, 108, 173.
(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, 29302962.
(14) De Feyter, S.; De Schryver, F. C. Two-Dimensional Supramolecular Self-Assembly Probed by Scanning Tunneling Microscopy.
Chem. Soc. Rev. 2003, 32, 139150.
(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, 722736.
(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, 702717.
(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,
27432752.
(20) Crassous, J. Chiral Transfer in Coordination Complexes:
Towards Molecular Materials. Chem. Soc. Rev. 2009, 38, 830840.
(21) Constable, E. C. Stereogenic Metal Centres-From Werner to
Supramolecular Chemistry. Chem. Soc. Rev. 2013, 42, 16371651.
(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, 25172534.
(24) Nonaka, K.; Yamaguchi, M.; Yasui, M. Guest-Induced Supramolecular Chirality in a Ditopic Azoprobe-Cyclodextrin Complex in
Water. Chem. Commun. 2014, 26, 1005910061.
(25) Joseph, G. Stereochemical Vocabulary for Structures that are
Chiral but not Asymmetric: History, Analysis, and Proposal for a
Rational Terminology. Chirality 2011, 23, 647659.
(26) Moss, G. P. Basic Terminology of Stereochemistry. Pure Appl.
Chem. 1996, 68, 21932222.
(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, 73047397

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, 1049810500.
(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,
55695571.
(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, 34083411.
(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, 81068112.
(36) Fischer, P.; Hache, F. Nonlinear Optical Spectroscopy of Chiral
Molecules. Chirality 2005, 17, 421437.
(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, 471485.
(38) Spitz, C.; Dahne, 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, 86648669.
(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, 50515056.
(40) Arteaga, O.; Canillas, A.; Purrello, R.; Ribo, J. M. Evidence of
Induced Chirality in Stirred Solutions of Supramolecular Nanofibers.
Opt. Lett. 2009, 34, 21772179.
(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, 45524554.
(42) Kunitake, T.; Kimizuka, N.; Higashi, N.; Nakashima, N. BilayerMembranes of Triple-Chain Ammonium Amphiphiles. J. Am. Chem. Soc.
1984, 106, 19781983.
(43) Nakashima, N.; Asakuma, S.; Kunitake, T. Optical Microscopic
Study of Helical Superstructuresof Chiral Bilayer Membranes. J. Am.
Chem. Soc. 1985, 107, 509510.
(44) Sorrenti, A.; Illa, O.; Ortuno, 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, 709713.
(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, 743750.
(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,
1000110010.
(48) Fuhrhop, J. H.; Helfrich, W. Fluid and Solid Fibers Made of Lipid
Molecular Bilayers. Chem. Rev. 1993, 93, 15651582.
(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, 14011443.
(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, 13121317.
(52) Li, Y.; Liu, M. Fabrication of Chiral Silver Nanoparticles and
Chiral Nanoparticulate Film via Organogel. Chem. Commun. 2008,
55715573.
(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, 80348040.
(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 QuinolinolGlutamide. Langmuir 2013, 29, 54355442.
(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, 33223329.
(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, 75017503.
(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, 92349241.
(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, 1607216079.
(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, 37023705.
(62) Wang, X.; Liu, M. Vicinal Solvent Effect on Supramolecular
Gelation: Alcohol Controlled Topochemical Reaction and the Toruloid
Nanostructure. Chem.Eur. J. 2014, 20, 1011010116.
(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, 74897494.
(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, 12741278.
(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, 587597.
(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, 39873990.
(67) Oda, R.; Huc, I.; Schmutz, M.; Candau, S. J.; MacKintosh, F. C.
Tuning Bilayer Twist Using Chiral Counterions. Nature 1999, 399,
566569.
(68) Berthier, D.; Buffeteau, T.; Leger, J.-M.; Oda, R.; Huc, I. From
Chiral Counterions to Twisted Membranes. J. Am. Chem. Soc. 2002, 124,
1348613494.
(69) Brizard, A.; Aime, 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, 37543762.
7386

DOI: 10.1021/cr500671p
Chem. Rev. 2015, 115, 73047397

Chemical Reviews

Review

(70) Fuhrhop, A. H.; Wang, T. Bolaamphiphiles. Chem. Rev. 2004, 104,


29012937.
(71) Shimizu, T.; Kogiso, M.; Masuda, M. Vesicle Assembly in
Microtubes. Nature 1996, 383, 487488.
(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, 462464.
(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, 1869418700.
(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, 47674769.
(75) Liu, Y.; Wang, T.; Huan, Y.; Li, Z.; He, G.; Liu, M. Self-Assembled
Supramolecular Nanotube Yarn. Adv. Mater. 2013, 25, 58755879.
(76) Masuda, M.; Shimizu, T. Lipid Nanotubes and Microtubes:
Experimental Evidence for Unsymmetrical Monolayer Membrane
Formation from Unsymmetrical Bolaamphiphiles. Langmuir 2004, 20,
59695977.
(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, 29462947.
(79) Moyer, T. J.; Cui, H.; Stupp, S. I. Tuning Nanostructure
Dimensions with Supramolecular Twisting. J. Phys. Chem. B 2013, 117,
46044610.
(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, 1246112468.
(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, 43834389.
(82) Soininen, A. J.; Kasemi, E.; Schlueter, 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, 1088210890.
(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, 15241527.
(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, 76627677.
(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, 51355151.
(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, 23112328.
(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, 63896395.
(88) Moberg, C. C3 Symmetry in Asymmetric Catalysis and Chiral
Recognition. Angew. Chem., Int. Ed. 1998, 37, 248268.
(89) Gibson, S. E.; Castaldi, M. P. Applications of Chiral C3-Symmetric
Molecules. Chem. Commun. 2006, 30453062.
(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, 292293.


(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, 10161022.
(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, 4246.
(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, 148155.
(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, 873876.
(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, 1475914769.
(96) Danila, I.; Riobe, F.; Piron, F.; Puigmarti-Luis, J.; Wallis, J. D.;
Linares, M.; gren, 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,
83448353.
(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.; Schon, 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, 253259.
(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, 43714373.
(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, 61256137.
(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, 810821.
(102) Aparicio, F.; Garca, F.; Sanchez, L. Supramolecular Polymerization of C3-Symmetric Organogelators: Cooperativity, Solvent, and
Gelation Relationship. Chem.Eur. J. 2013, 19, 32393248.
(103) Garca, F.; Buendia, J.; Sanchez, L. Supramolecular Ribbons from
Amphiphilic Trisamides Self-Assembly. J. Org. Chem. 2011, 76, 6271
6276.
(104) Garca, F.; Sanchez, 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,
734742.
(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, 1176111770.
(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, 55465550.
(107) Wang, C.; Dong, H.; Hu, W.; Liu, Y.; Zhu, D. Semiconducting Conjugated Systems in Field-Effect Transistors: A Material Odyssey of
Organic Electronics. Chem. Rev. 2012, 112, 22082267.
(108) Kim, F. S.; Ren, G.; Jenekhe, S. A. One-Dimensional
Nanostructures of -Conjugated Molecular Systems: Assembly, Proper7387

DOI: 10.1021/cr500671p
Chem. Rev. 2015, 115, 73047397

Chemical Reviews

Review

ties, and Applications from Photovoltaics, Sensors, and Nanophotonics


to Nanoelectronics. Chem. Mater. 2011, 23, 682732.
(109) Beaujuge, P. M.; Frechet, J. M. Molecular Design and Ordering
Effects in -Functional Materials for Transistor and Solar Cell
Applications. J. Am. Chem. Soc. 2011, 133, 2000920029.
(110) Gorl, D.; Zhang, X.; Wurthner, F. Molecular Assemblies of
Perylene Bisimide Dyes in Water. Angew. Chem., Int. Ed. 2012, 51,
63286348.
(111) Babu, S. S.; Praveen, V. K.; Ajayaghosh, A. Functional piGelators and Their Applications. Chem. Rev. 2014, 114, 19732129.
(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, 42224242.
(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, 1080110806.
(115) Xie, Z.; Stepanenko, V.; Radacki, K.; Wurthner, F. Chiral JAggregates of Atropo-Enantiomeric Perylene Bisimides and Their SelfSorting Behavior. Chem.Eur. J. 2012, 18, 70607070.
(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, 56445653.
(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, 75767584.
(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, 1087110876.
(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, 316321.
(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, 81818186.
(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, 15211526.
(123) Aparicio, F.; Nieto-Ortega, B.; Najera, F.; Ramrez, F. J.;
Navarrete, J. T. L.; Casado, J.; Sanchez, L. Inversion of Supramolecular
Helicity in Oligo-p-phenylene-Based Supramolecular Polymers: Influence of Molecular Atropisomerism. Angew. Chem., Int. Ed. 2014, 53,
13731377.
(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, 39393942.
(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, 15281538.
(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,
26482651.
(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, 80578065.
(129) Narayan, B.; Senanayak, S. P.; Jain, A.; Narayan, K. S.; George, S.
J. Self-Assembly of -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, 1457614583.
(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, 38473854.
(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, 16141623.
(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, 20512060.
(134) Henze, O.; Feast, W. J.; Gardebien, F.; Jonkheijm, P.; Lazzaroni,
R.; Leclere, P.; Meijer, E. W.; Schenning, A. Chiral Amphiphilic SelfAssembled --Linked Quinque-, Sexi-, and Septithiophenes: Synthesis, Stability and Odd-Even Effects. J. Am. Chem. Soc. 2006, 128,
59235929.
(135) Digennaro, A.; Wennemers, H.; Joshi, G.; Schmid, S.; MenaOsteritz, E.; Bauerle, P. Chiral Suprastructures of Asymmetric
Oligothiophene-Hybrids Induced by a Single Proline. Chem. Commun.
2013, 49, 1092910931.
(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, 45114521.
(137) Nilsson, K. P. R.; Rydberg, J.; Baltzer, L.; Inganas, 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, 1119711202.
(138) Van den Bergh, K.; Cosemans, I.; Verbiest, T.; Koeckelberghs, G.
Expression of Supramolecular Chirality in Block Copoly(thiophene)s.
Macromolecules 2010, 43, 37943800.
(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 PtPt and pi-pi
Interactions. Chem.Eur. J. 2009, 15, 47754778.
(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, 79867991.
(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,
48184822.
(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, 60136017.
(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, 16881692.
7388

DOI: 10.1021/cr500671p
Chem. Rev. 2015, 115, 73047397

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, 34513453.
(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, 10921098.
(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, 1778917796.
(147) Samanta, S. K.; Bhattacharya, S. Excellent Chirality Transcription in Two-Component Photochromic Organogels Assembled
through J-Aggregation. Chem. Commun. 2013, 49, 14251427.
(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, 70917099.
(149) Zhang, W.; Fujiki, M.; Zhu, X. Chiroptical Nanofibers Generated
from Achiral Metallophthalocyanines Induced by Diamine Homochirality. Chem.Eur. J. 2011, 17, 1062810635.
(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, 34293437.
(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, 1776417771.
(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, 98119819.
(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, 1190311906.
(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, 1468014688.
(155) Watanabe, K.; Iida, H.; Akagi, K. Circularly Polarized Blue
Luminescent Spherulites Consisting of Hierarchically Assembled Ionic
Conjugated Polymers with a Helically -Stacked Structure. Adv. Mater.
2012, 24, 64516456.
(156) Kumar, M.; George, S. J. Homotropic And Heterotropic
Allosteric Regulation of Supramolecular Chirality. Chem. Sci. 2014, 5,
30253030.
(157) Escarcega-Bobadilla, M. V.; Salassa, G.; Belmonte, M. M.;
Escudero-Adan, E. C.; Kleij, A. W. Versatile Switching in Substrate
Topicity: Supramolecular Chirality Induction in Di- and Trinuclear
Host Complexes. Chem.Eur. J. 2012, 18, 68056810.
(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, 49164922.
(159) Rivera, J. M. Chiral Spaces: Dissymmetric Capsules through SelfAssembly. Science 1998, 279, 10211023.
(160) Rivera, J. M.; Craig, S. L.; Martn, T.; Rebek, J. Chiral Guests and
Their Ghosts in Reversibly Assembled Hosts. Angew. Chem., Int. Ed.
2000, 39, 21302132.
(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,
72437247.

(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, 28322838.
(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, 2019720202.
(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, 89348936.
(165) Bosnich, B. Asymmetric Syntheses Asymmetric Transformations and Asymmetric Inductions in an Optically Active Solvent. J. Am.
Chem. Soc. 1967, 89, 61436148.
(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, 49414942.
(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, 69916995.
(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, 48474848.
(169) Stepanenko, V.; Li, X.; Gershberg, J.; Wurthner, F. Evidence for
Kinetic Nucleation in Helical Nanofiber Formation Directed by Chiral
Solvent for a Perylene Bisimide Organogelator. Chem.Eur. J. 2013, 19,
41764183.
(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, 66366638.
(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, 51055111.
(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, 173177.
(174) Yashima, E.; Maeda, K.; Iida, H.; Furusho, Y.; Nagai, K. Helical
Polymers: Synthesis, Structures, and Functions. Chem. Rev. 2009, 109,
61026211.
(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, 54955503.
(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, 33453350.
(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, 43294342.
(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, 73047397

Chemical Reviews

Review

Columnar Assemblies with Phototunable Chirality. J. Am. Chem. Soc.


2012, 22, 1802518032.
(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, 1149211498.
(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-Urena, F.; LopezGonzalez, J. J. Photoinduced Chiral Nematic Organization in an Achiral
Glassy Nematic Azopolymer. Adv. Funct. Mater. 2007, 17, 34863492.
(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, 37583765.
(203) Gopal, A.; Hifsudheen, M.; Furumi, S.; Takeuchi, M.;
Ajayaghosh, A. Thermally Assisted Photonic Inversion of Supramolecular Handedness. Angew. Chem., Int. Ed. 2012, 51, 1050510509.
(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,
1820518208.
(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, 1015010159.
(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, 1465014659.
(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, 1095410956.
(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, 81118115.
(210) Kumar, M.; Jonnalagadda, N.; George, S. J. Molecular
Recognition Driven Self-Assembly and Chiral Induction in Naphthalene
Diimide Amphiphiles. Chem. Commun. 2012, 48, 1094810950.
(211) Ke, D.; Tang, A.; Zhan, C.; Yao, J. Conformation-Variable PDI@
Beta-Sheet Nanohelices Show Stimulus-Responsive Supramolecular
Chirality. Chem. Commun. 2013, 49, 49144916.
(212) Maity, S.; Kumar, P.; Haldar, D. Sonication-Induced Instant
Amyloid-Like Fibril Formation and Organogelation by a Tripeptide. Soft
Matter 2011, 7, 52395245.
(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, 79607966.
(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, 90999106.
(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, 44944498.
(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, 60306037.

Polymer Backbone Stimulated by the Pendant Conformational Change.


Chem.Eur. J. 2004, 10, 40004010.
(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, 1171111717.
(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, 60036008.
(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, 76397650.
(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, 10041013.
(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, 74127423.
(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, 20222027.
(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, 85408543.
(188) Zhang, L.; Liu, M. Supramolecular Chirality and Chiral Inversion
of Tetraphenylsulfonato Porphyrin Assemblies on Optically Active
Polylysine. J. Phys. Chem. B 2009, 113, 1401514020.
(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, 76157622.
(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,
55545556.
(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, 41494158.
(192) Duan, P.; Cao, H.; Zhang, L.; Liu, M. Gelation Induced
Supramolecular Chirality: Chirality Transfer, Amplification and
Application. Soft Matter 2014, 10, 54285448.
(193) Gutz, C.; Hovorka, R.; Struch, N.; Bunzen, J.; Meyer-Eppler, G.;
Qu, Z.; Grimme, S.; Topc, F.; Rissanen, K.; Cetina, M.; Engeser, M.;
Lutzen, A. Enantiomerically Pure Trinuclear Helicates via Diastereoselective Self-Assembly and Characterization of Their Redox
Chemistry. J. Am. Chem. Soc. 2014, 136, 1183011838.
(194) Adhikari, B.; Kraatz, H. B. Redox-Triggered Changes in the SelfAssembly of a Ferrocene-Peptide Conjugate. Chem. Commun. 2014, 50,
55515553.
(195) del Barrio, J.; Tejedor, R. M.; Chinelatto, L. S.; Sanchez, C.;
Pinol, M.; Oriol, L. Photocontrol of the Supramolecular Chirality
Imposed by Stereocenters in Liquid Crystalline Azodendrimers. Chem.
Mater. 2010, 22, 17141723.
(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, 17781782.
(197) Tejedor, R. M.; Serrano, J. L.; Oriol, L. Photocontrol of
Supramolecular Architecture in Azopolymers: Achiral and Chiral
Aggregation. Eur. Polym. J. 2009, 45, 25642571.
(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, 73047397

Chemical Reviews

Review

(238) Lauceri, R.; DUrso, 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, 1155411559.
(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, 47974805.
(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, 954962.
(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, 13941403.
(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, 1223812246.
(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, 1675316755.
(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, 937944.
(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,
114117.
(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, 1799918006.
(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, 63506355.
(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, 20632066.
(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, 12591268.
(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, 36763682.
(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,
510516.
(254) Young, W. R.; Aviram, A.; Cox, R. J. Stilbene Derivatives. New
Class of Room Temperature Nematic Liquids. J. Am. Chem. Soc. 1972,
94, 39763981.
(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, 30513066.
(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,
32293232.

(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, 28372839.
(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,
24202427.
(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, 23022312.
(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, 88508858.
(222) Palmans, A. R.; Meijer, E. W. Amplification of Chirality in
Dynamic Supramolecular Aggregates. Angew. Chem., Int. Ed. 2007, 46,
89488968.
(223) Kulkarni, C.; Munirathinam, R.; George, S. J. Self-Assembly of
Coronene Bisimides: Mechanistic Insight and Chiral Amplification.
Chem.Eur. J. 2013, 19, 1127011278.
(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, 620626.
(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, 82538261.
(226) van Gestel, J. Amplification of Chirality in Helical Supramolecular Polymers: The Majority-Rules Principle. Macromolecules
2004, 37, 38943898.
(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, 754758.
(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, 35983608.
(229) Aparicio, F.; Vicente, F.; Sanchez, L. Amplification of Chirality in
N,N-1,2-Ethanediylbisbenzamides: From Planar Sheets to Twisted
Ribbons. Chem. Commun. 2010, 46, 83568358.
(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, 429434.
(232) Takashima, S.; Abe, H.; Inouye, M. Copper(II)/PhenanthrolineMediated CD-Enhancement and Chiral Memory Effect on a MetaEthynylpyridine Oligomer. Chem. Commun. 2012, 48, 33303332.
(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, 1235312354.
(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, 1090010908.
(235) Rosaria, L.; DUrso, A.; Mammana, A.; Purrello, R. Chiral
Memory: Induction, Amplification, and Switching in Porphyrin
Assemblies. Chirality 2008, 20, 411419.
(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, 51595163.
(237) Mammana, A.; DUrso, A.; Lauceri, R.; Purrello, R. Switching Off
and On the Supramolecular Chiral Memory in Porphyrin Assemblies. J.
Am. Chem. Soc. 2007, 129, 80628063.
7391

DOI: 10.1021/cr500671p
Chem. Rev. 2015, 115, 73047397

Chemical Reviews

Review

(258) Hough, L. E.; Spannuth, M.; Nakata, M.; Coleman, D. A.; Jones,
C. D.; Dantlgraber, G.; Tschierske, C.; Watanabe, J.; Korblova, E.;
Walba, D. M.; Maclennan, J. E.; Glaser, M. A.; Clark, N. A. Chiral
Isotropic Liquids from Achiral Molecules. Science 2009, 325, 452456.
(259) Hough, L. E.; Jung, H. T.; Kruerke, D.; Heberling, M. S.; Nakata,
M.; Jones, C. D.; Chen, D.; Link, D. R.; Zasadzinski, J.; Heppke, G.;
Rabe, J. P.; Stocker, W.; Korblova, E.; Walba, D. M.; Glaser, M. A.; Clark,
N. A. Helical Nanofilament Phases. Science 2009, 325, 456460.
(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, 1236812372.
(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, 68636866.
(262) Zhang, C.; Diorio, N.; Lavrentovich, O. D.; Jakli, 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, 72987332.
(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, 27692774.
(266) Takeuchi, M.; Tanaka, S.; Shinkai, S. On The Influence of
Porphyrin - 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, 20732082.
(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, 60406043.
(269) DeRossi, U.; Dahne, 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, 760763.
(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, 96449652.
(271) Mineo, P.; Villari, V.; Scamporrino, E.; Micali, N. Supramolecular Chirality Induced by A Weak Thermal Force. Soft Matter
2014, 10, 4447.
(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,
1129711301.
(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, 4043.
(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, 64026409.
(275) Terech, P.; Weiss, R. G. Low Molecular Mass Gelators of
Organic Liquids and The Properties of Their Gels. Chem. Rev. 1997, 97,
31333160.
(276) Hirst, A. R.; Smith, D. K. Two-Component Gel-Phase MaterialsHighly Tunable Self-Assembling Systems. Chem.Eur. J. 2005, 11,
54965508.
(277) Babu, S. S.; Praveen, V. K.; Ajayaghosh, A. Functional -Gelators
and Their Applications. Chem. Rev. 2014, 114, 19732129.

(278) Estroff, L. A.; Hamilton, A. D. Water Gelation by Small Organic


Molecules. Chem. Rev. 2004, 104, 12011218.
(279) Yang, Z.; Liang, G.; Xu, B. Enzymatic Hydrogelation of Small
Molecules. Acc. Chem. Res. 2008, 41, 315326.
(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, 70117015.
(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, 493499.
(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, 57325738.
(283) Zhang, S.; Yang, S.; Lan, J.; Yang, S.; You, J. Helical Nonracemic
Tubular Coordination Polymer Gelators from Simple Achiral
Molecules. Chem. Commun. 2008, 61706172.
(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,
1342413428.
(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 LangmuirBlodgett Films of An Achiral Amphiphilic
Barbituric Acid. J. Am. Chem. Soc. 2004, 126, 13221323.
(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, 41554162.
(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,
18241829.
(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,
25652569.
(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, 722729.
(292) Huang, X.; Liu, M. Chirality of Photopolymerized Organized
Supramolecular Polydiacetylene Films. Chem. Commun. 2003, 6667.
(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, 22292235.
(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, 47804784.
(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, 40884098.
(296) Wang, T.; Liu, M. Langmuir-Schaefer Films of A Set of Achiral
Amphiphilic Porphyrins: Aggregation and Supramolecular Chirality.
Soft Matter 2008, 4, 775783.
(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, 73047397

Chemical Reviews

Review

Achiral Porphyrins at The Air/Aqueous Interface by Second Harmonic


Generation Linear Dichroism. J. Phys. Chem. C 2014, 118, 67266733.
(298) Crusats, J.; El-Hachemi, Z.; Ribo, J. M. Hydrodynamic Effects on
Chiral Induction. Chem. Soc. Rev. 2010, 39, 569577.
(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, 80328035.
(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, 81988202.
(301) DUrso, A.; Randazzo, R.; LoFaro, L.; Purrello, R. Vortexes and
Nanoscale Chirality. Angew. Chem., Int. Ed. 2010, 49, 108112.
(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, 92889292.
(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,
1247412477.
(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, 88208826.
(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,
201207.
(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.; Menard, F.; Tamura, M. Circular Polarization in
Star-Formation Regions: Implications for Biomolecular Homochirality.
Science 1998, 281, 672674.
(308) Cronin, J. R.; Pizzarello, S. Enantiomeric Excesses in Meteoritic
Amino Acids. Science 1997, 275, 951955.
(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, 45184522.
(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, 365367.
(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, 1210812114.
(313) Safont-Sempere, M. M.; Fernandez, G.; Wurthner, F. SelfSorting Phenomena in Complex Supramolecular Systems. Chem. Rev.
2011, 111, 57845814.
(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, 76707678.
(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, 1396213965.
(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, 93209322.
(318) Safont-Sempere, M. M.; Osswald, P.; Stolte, M.; Grune, M.;
Renz, M.; Kaupp, M.; Radacki, K.; Braunschweig, H.; Wurthner, F.
Impact of Molecular Flexibility on Binding Strength and Self-Sorting of
Chiral -Surfaces. J. Am. Chem. Soc. 2011, 133, 95809591.
(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, 71697185.
(320) Makarevic, J.; Jokic, M.; Raza, Z.; Stefanic, Z.; Kojic-Prodic, B.;
Z inic, M. Chiral Bis(Amino Alcohol)Oxalamide GelatorsGelation
Properties and Supramolecular Organization: Racemate versus Pure
Enantiomer Gelation. Chem.Eur. J. 2003, 9, 55675580.
(321) Amemiya, R.; Mizutani, M.; Yamaguchi, M. Two-Component
Gel Formation by Pseudoenantiomeric Ethynylhelicene Oligomers.
Angew. Chem., Int. Ed. 2010, 49, 19951999.
(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, 1497514977.
(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, 1077210778.
(324) Lee, S. J.; Lin, W. Chiral Metallocycles: Rational Synthesis and
Novel Applications. Acc. Chem. Res. 2008, 41, 521537.
(325) Pescitelli, G.; Di Bari, L.; Berova, N. Application of Electronic
Circular Dichroism in The Study of Supramolecular Systems. Chem. Soc.
Rev. 2014, 43, 52115233.
(326) Borovkov, V. Supramolecular Chirality in Porphyrin Chemistry.
Symmetry 2014, 6, 256294.
(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)32(alpha-CDX). Chem. Commun. 2010, 46, 35353537.
(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, 21442152.
(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, 20332045.
(330) Berova, N.; Pescitelli, G.; Petrovic, A. G.; Proni, G. Probing
Molecular Chirality by CD-Sensitive Dimeric Metalloporphyrin Hosts.
Chem. Commun. 2009, 59585980.
(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, 129139.
(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, 449459.
(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, 349358.
(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, 22122221.
(336) Zhang, D.-W.; Zhao, X.; Li, Z.-T. Aromatic Amide and
Hydrazide Foldamer-Based Responsive Host-Guest Systems. Acc.
Chem. Res. 2014, 47, 19611970.
7393

DOI: 10.1021/cr500671p
Chem. Rev. 2015, 115, 73047397

Chemical Reviews

Review

(357) Edwards, W.; Smith, D. K. Enantioselective Component


Selection in Multicomponent Supramolecular Gels. J. Am. Chem. Soc.
2014, 136, 11161124.
(358) Zhang, X.; Yin, J.; Yoon, J. Recent Advances in Development of
Chiral Fluorescent and Colorimetric Sensors. Chem. Rev. 2014, 114,
49184959.
(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,
1220012203.
(361) Wolf, C.; Bentley, K. W. Chirality Sensing Using Stereodynamic
Probes with Distinct Electronic Circular Dichroism Output. Chem. Soc.
Rev. 2013, 42, 54085424.
(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, 216224.
(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, 1032010335.
(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 HostGuest Complexes with Chiral
Alcohol and Monoamine Conjugates. J. Am. Chem. Soc. 2001, 123,
59745982.
(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, 59625973.
(367) Proni, G.; Pescitelli, G.; Huang, X.; Nakanishi, K.; Berova, N.
Magnesium Tetraarylporphyrin Tweezer: A CD-Sensitive Host for
Absolute Configurational Assignments of -Chiral Carboxylic Acids. J.
Am. Chem. Soc. 2003, 125, 1291412927.
(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, 18851893.
(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, 1689616904.
(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, 69006902.
(371) Feng, L.; Xu, B.; Ren, J.; Zhao, C.; Qu, X. A Human Telomeric
DNA-Based Chiral Biosensor. Chem. Commun. 2012, 48, 90689070.
(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,
20312035.
(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, 943948.
(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,
56945699.

(337) Mutihac, L.; Lee, J. H.; Kim, J. S.; Vicens, J. Recognition of


Amino Acids by Functionalized Calixarenes. Chem. Soc. Rev. 2011, 40,
27772779.
(338) Harada, A. Cyclodextrin-Based Molecular Machines. Acc. Chem.
Res. 2001, 34, 456464.
(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, 337347.
(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, 19821994.
(341) Ghale, G.; Nau, W. M. Dynamically Analyte-Responsive
Macrocyclic Host-Fuorophore Systems. Acc. Chem. Res. 2014, 47,
21502159.
(342) Liu, Y.; Chen, Y. Cooperative Binding and Multiple Recognition
by Bridged Bis(-Cyclodextrin)s with Functional Linkers. Acc. Chem.
Res. 2006, 39, 681691.
(343) Ma, X.; Tian, H. Stimuli-Responsive Supramolecular Polymers
in Aqueous Solution. Acc. Chem. Res. 2014, 47, 19711981.
(344) Wang, M.-X. Nitrogen and Oxygen Bridged Calixaromatics:
Synthesis, Structure, Functionalization, and Molecular Recognition. Acc.
Chem. Res. 2012, 45, 182195.
(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, 19952005.
(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, 27562781.
(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, 25952601.
(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, 21442152.
(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, 1487114880.
(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, 32173226.
(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, 1548515488.
(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, 72977299.
(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, 66016605.
(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, 48814883.
7394

DOI: 10.1021/cr500671p
Chem. Rev. 2015, 115, 73047397

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, 1196611976.
(376) Rodrigues, S. P.; Lan, S.; Kang, L.; Cui, Y.; Cai, W. Nonlinear
Imaging and Spectroscopy of Chiral Metamaterials. Adv. Mater. 2014,
26, 61576162.
(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, 49914997.
(379) Cui, Y.; Kang, L.; Lan, S.; Rodrigues, S.; Cai, W. Giant Chiral
Optical Response from a Twisted-Arc Metamaterial. Nano Lett. 2014,
14, 10211025.
(380) He, Y.; Larsen, G. K.; Ingram, W.; Zhao, Y. Tunable ThreeDimensional Helically Stacked Plasmonic Layers on Nanosphere
Monolayers. Nano Lett. 2014, 14, 19761981.
(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,
783787.
(382) Qiu, Y.; Chen, P.; Guo, P.; Li, Y.; Liu, M. Supramolecular
Chiroptical Switches Based on Achiral Molecules. Adv. Mater. 2008, 20,
29082913.
(383) Hecht, S. Optical Switching of Hierarchical Self-Assembly:
Towards Enlightened Materials. Small 2005, 1, 2629.
(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 121179.
(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, 45414552.
(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, 1393813941.
(387) Guo, P.; Zhang, L.; Liu, M. A Supramolecular Chiroptical Switch
Exclusively from an Achiral Amphiphile. Adv. Mater. 2006, 18, 177180.
(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, 285291.
(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, 13261331.
(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, 59095917.
(391) Kim, M.-J.; Yoo, S.-J.; Kim, D.-Y. A Supramolecular Chiroptical
Switch Using an Amorphous Azobenzene Polymer. Adv. Funct. Mater.
2006, 16, 20892094.
(392) Canary, J. W. Redox-Triggered Chiroptical Molecular Switches.
Chem. Soc. Rev. 2009, 38, 747756.
(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, 1787917884.
(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, 1562815631.

(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, 1209912101.
(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, 1082610829.
(397) Goto, H.; Yashima, E. Electron-Induced Switching of The
Supramolecular Chirality of Optically Active Polythiophene Aggregates.
J. Am. Chem. Soc. 2002, 124, 79437949.
(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, 1431214313.
(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, 1507615086.
(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, 36933696.
(401) Wang, J.; Feringa, B. L. Dynamic Control of Chiral Space in a
Catalytic Asymmetric Reaction Using a Molecular Motor. Science 2011,
331, 14291432.
(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, 1120911221.
(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, 1691116921.
(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, 46544660.
(405) Cesar, V.; Bellemin-Laponnaz, S.; Gade, L. H. Chiral NHeterocyclic Carbenes as Stereodirecting Ligands in Asymmetric
Catalysis. Chem. Soc. Rev. 2004, 33, 619636.
(406) Ding, K.; Wang, Z.; Wang, X.; Liang, Y.; Wang, X. SelfSupported Chiral Catalysts for Heterogeneous Enantioselective
Reactions. Chem.Eur. J. 2006, 12, 51885197.
(407) Li, C.; Zhang, H.; Jiang, D.; Yang, Q. Chiral Catalysis in
Nanopores of Mesoporous Materials. Chem. Commun. 2007, 547548.
(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, 11841195.
(409) Wohlgemuth, R. Asymmetric Biocatalysis with Microbial
Enzymes and Cells. Curr. Opin. Microbiol. 2010, 13, 283292.
(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, 41264139.
(411) Lee, Y. S.; Alam, M. M.; Keri, R. S. Enantioselective Reactions of
N-Acyliminium Ions Using Chiral Organocatalysts. Chem. Asian. J.
2013, 8, 29062919.
(412) Ding, K. Synergistic Effect of Binary Component Ligands in
Chiral Catalyst Library Engineering for Enantioselective Reactions.
Chem. Commun. 2008, 909921.
(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, 3444.
(414) Palomo, C.; Oiarbide, M.; Garcia, J. M. Current Progress in The
Asymmetric Aldol Addition Reaction. Chem. Soc. Rev. 2004, 33, 6575.
(415) Ouellet, S. G.; Walji, A. M.; MacMillan, D. W. C.
Enantioselective Organocatalytic Transfer Hydrogenation Reactions
Using Hantzsch Esters. Acc. Chem. Res. 2007, 40, 13271339.
(416) Mlynarski, J.; Paradowska, J. Catalytic Asymmetric Aldol
Reactions in Aqueous Media. Chem. Soc. Rev. 2008, 37, 15021511.
7395

DOI: 10.1021/cr500671p
Chem. Rev. 2015, 115, 73047397

Chemical Reviews

Review

(417) Yoon, M.; Srirambalaji, R.; Kim, K. Homochiral Metal-Organic


Frameworks for Asymmetric Heterogeneous Catalysis. Chem. Rev. 2011,
112, 11961231.
(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, 66756677.
(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, 72167228.
(421) Yang, C.; Inoue, Y. Supramolecular Photochirogenesis. In
Supramolecular Photochemistry; John Wiley & Sons, Inc.: New York,
2011; pp 115153.
(422) Yang, C.; Inoue, Y. Supramolecular Photochirogenesis. Chem.
Soc. Rev. 2014, 43, 41234143.
(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,
36273630.
(425) Uraguchi, D.; Ueki, Y.; Ooi, T. Chiral Organic Ion Pair Catalysts
Assembled through a Hydrogen-Bonding Network. Science 2009, 326,
120123.
(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,
1218912192.
(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, 10551059.
(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, 1384713853.
(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, 77617765.
(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,
47424745.
(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, 1510315113.
(432) Roelfes, G.; Feringa, B. L. DNA-Based Asymmetric Catalysis.
Angew. Chem., Int. Ed. 2005, 44, 32303232.
(433) Coquiere, D.; Feringa, B. L.; Roelfes, G. DNA-Based Catalytic
Enantioselective Michael Reactions in Water. Angew. Chem., Int. Ed.
2007, 46, 93089311.
(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, 33463348.
(435) Oelerich, J.; Roelfes, G. DNA-Based Asymmetric Organometallic Catalysis in Water. Chem. Sci. 2013, 4, 20132017.
(436) Boersma, A. J.; Coquiere, 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, 991995.

(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, 43094311.
(438) Wang, C.; Jia, G.; Zhou, J.; Li, Y.; Liu, Y.; Lu, S.; Li, C.
Enantioselective DielsAlder Reactions with G-Quadruplex DNABased Catalysts. Angew. Chem., Int. Ed. 2012, 51, 93529355.
(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, 1154611549.
(440) Pace, T. C. S.; Muller, V.; Li, S.; Lincoln, P.; Andreasson, J.
Enantioselective Cyclization of Photochromic Dithienylethenes Bound
to DNA. Angew. Chem., Int. Ed. 2013, 52, 43934396.
(441) Coquiere, D.; Bos, J.; Beld, J.; Roelfes, G. Enantioselective
Artificial Metalloenzymes Based on a Bovine Pancreatic Polypeptide
Scaffold. Angew. Chem., Int. Ed. 2009, 48, 51595162.
(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, 75997603.
(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, 198.
(444) Ben-Moshe, V.; Beratan, D. N.; Nitzan, A.; Skourtis, S. S. Chiral
Control of Current Transfer in Molecules. Top. Curr. Chem. 2011, 298,
259278.
(445) Mori, T.; Inoue, Y. Recent Theoretical and Experimental
Advances in The Electronic Circular Dichroisms of Planar Chiral
Cyclophanes. Top. Curr. Chem. 2011, 298, 99128.
(446) Lee, B.; Lee, S.-Y. Polarization-Dependent Plasmonic Chiral
Devices in Nanoplasmonics 113135, 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, 47734775.
(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, 1089110896.
(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, 578581.
(450) Yang, Y.; Zhang, Y.; Wei, Z. Supramolecular Helices: Chirality
Transfer from Conjugated Molecules to Structures. Adv. Mater. 2013,
25, 60396049.
(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, 913915.
(452) Rethore, C.; Avarvari, N.; Canadell, E.; Auban-Senzier, P.;
Fourmigue, M. Chiral Molecular Metals: Syntheses, Structures, and
Properties of The AsF6-Salts of Racemic ()-, (R)-, and (S)TetrathiafulvaleneOxazoline Derivatives. J. Am. Chem. Soc. 2005,
127, 57485749.
(453) Yan, Y.; Wang, R.; Qiu, X.; Wei, Z. Hexagonal Superlattice of
Chiral Conducting Polymers Self-Assembled by Mimicking -Sheet
Proteins with Anisotropic Electrical Transport. J. Am. Chem. Soc. 2010,
132, 1200612012.
(454) Feng, X.; Marcon, V.; Pisula, W.; Hansen, M. R.; Kirkpatrick, J.;
Grozema, F.; Andrienko, D.; Kremer, K.; Mullen, K. Towards High
Charge-Carrier Mobilities by Rational Design of The Shape and
Periphery of Discotics. Nat. Mater. 2009, 8, 421426.
(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,
27562757.
7396

DOI: 10.1021/cr500671p
Chem. Rev. 2015, 115, 73047397

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, 634638.
(457) Mislow, K. Absolute Asymmetric Synthesis: a Commentary.
Collect. Czech. Chem. Commun. 2003, 68, 849864.
(458) Carr, R.; Evans, N. H.; Parker, D. Lanthanide Complexes as
Chiral Probes Exploiting Circularly Polarized Luminescence. Chem. Soc.
Rev. 2012, 41, 76737686.
(459) Bunzli, J.-C. G.; Piguet, C. Taking Advantage of Luminescent
Lanthanide Ions. Chem. Soc. Rev. 2005, 34, 10481077.
(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, 99099910.
(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,
83378347.
(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, 68306831.
(463) Satrijo, A.; Meskers, S. C. J.; Swager, T. M. Probing a Conjugated
Polymers Transfer of Organization-Dependent Properties from
Solutions to Films. J. Am. Chem. Soc. 2006, 128, 90309031.
(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, 43224328.
(465) Watanabe, K.; Osaka, I.; Yorozuya, S.; Akagi, K. Helically Stacked 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, 1189911907.
(467) Field, J. E.; Muller, G.; Riehl, J. P.; Venkataraman, D. Circularly
Polarized Luminescence from Bridged Triarylamine Helicenes. J. Am.
Chem. Soc. 2003, 125, 1180811809.
(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 -Conjugated Molecules. J. Am. Chem. Soc. 2011, 133, 92669269.
(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 ReceptorAnion 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, 1409014097.
(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 HydrogelEnzyme Hybrids.
Nat. Chem. 2014, 6, 511518.
(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, 99079914.
(473) McClendon, M. T.; Stupp, S. I. Tubular Hydrogels of
Circumferentially Aligned Nanofibers to Encapsulate and Orient
Vascular Cells. Biomaterials 2012, 33, 57135722.
(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,
84198422.

(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, 51725180.
(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, 77897793.
(477) Das, R. K.; Zouani, O. F.; Labrugere, C.; Oda, R.; Durrieu, M.-C.
Influence of Nanohelical Shape and Periodicity on Stem Cell Fate. ACS
Nano 2013, 7, 33513361.
(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, 1887818884.

7397

DOI: 10.1021/cr500671p
Chem. Rev. 2015, 115, 73047397

Anda mungkin juga menyukai