You are on page 1of 70

Review

pubs.acs.org/CR

Tetrathiafulvalene- (TTF‑) Derived Oligopyrrolic Macrocycles


Atanu Jana,†,‡ Masatoshi Ishida,§ Jung Su Park,∥ Steffen Baḧ ring,⊥ Jan O. Jeppesen,*,⊥
and Jonathan L. Sessler*,‡,#

Department of Chemistry, University of Sheffield, Sheffield S10 2TN, United Kingdom
§
Department of Chemistry and Biochemistry, Graduate School of Engineering and Center for Molecular Systems, Kyushu University,
Fukuoka 819-0395, Japan

Department of Chemistry, Sookmyung Womens’s University, Seoul 140-742, South Korea

Department of Physics, Chemistry, and Pharmacy, University of Southern Denmark, Campusvej 55, 5230, Odense M, Denmark
#
Department of Chemistry, The University of Texas at Austin, Austin, Texas 78712-1224, United States

Institute for Supramolecular Chemistry and Catalysis, Shanghai University, Shanghai, 200444, China

ABSTRACT: After the epochal discovery of the “organic metal”, namely, tetrathiafulva-
lene (TTF)−7,7,8,8-tetracyano-p-quinodimethane (TCNQ) dyad in 1973, scientists have
made efforts to derivatize TTF for constructing various supramolecular architectures to
control the charge-transfer processes by adjusting the donor−acceptor strength of the
dyads for numerous applications. The interesting inherent electronic donor properties of
TTFs control the overall electrochemical properties of the supramolecular structures,
leading to the construction of highly efficient optoelectronic materials, photovoltaic solar
cells, organic field-effect transistors, and optical sensors. Modified TTF structures thus
constitute promising candidates for the development of so-called “functional materials” that
could see use in modern technological applications. The versatility of the TTF unit and the
pioneering synthetic strategies that have been developed over the past few decades provide
opportunities to tune the architecture and function for specific purposes. This review
covers the “state of the art” associated with TTF-annulated oligopyrrolic macrocyclic
compounds. Points of emphasis include synthesis, properties, and potential applications.

CONTENTS 2.2.1. “Pacman”-Type Schiff-Base TTF-Calix[4]-


pyrrole 2663
1. Introduction 2642 2.2.2. TTF-Annulated Calix[2]pyrrole[2]-
2. TTF-Annulated Oligopyrrolic Macrocyclic Com- thiophene 2664
pounds 2646 2.3. TTF-Annulated Expanded Calix[n]pyrroles 2664
2.1. TTF-Annulated Calix[4]pyrroles 2646 2.4. TTF-Annulated Porphyrins 2665
2.1.1. TTF-Annulated Calix[4]pyrroles for Anion 2.4.1. TTF-Porphyrins with Different Numbers
Sensing 2647 of Fused TTF Units 2665
2.1.2. TTF-Annulated Calix[4]pyrroles for Sens- 2.4.2. Quinoxaline-Fused TTF-Porphyrins 2669
ing Neutral Aromatic Electron-Deficient 2.4.3. TTF-Annulated Expanded Porphyrins 2670
Guests 2649 2.5. ExTTF Porphyrins 2671
2.1.3. TTF-Annulated Calix[4]pyrroles for Sens- 2.6. Other Porphyrin−TTF-Based Donor−Accept-
ing Spherical Guests 2653 or Systems 2673
2.1.4. TTF-Annulated Calix[4]pyrroles and Sur- 2.6.1. Donor−Acceptor Systems Based on
face Studies 2656 Porphyrins with Peripheral TTF Substitu-
2.1.5. TTF-Annulated Calix[4]pyrroles for Che- ents Connected through Flexible Organ-
moresponsive Supramolecular Copoly- ic Spacers 2673
merization with Heterocomplementary 2.6.2. Donor−Acceptor Systems Consisting of
Calix[4]pyrroles 2657 Porphyrins with Axially Coordinated TTF
2.1.6. TTF-Annulated Calix[4]pyrroles for Ion- Derivatives 2677
Mediated Reversible Electron-Transfer
(ET) Process in Supramolecular Ensem-
bles Formed with Electron-Deficient
Guests 2661 Special Issue: Expanded, Contracted, and Isomeric Porphyrins
2.2. TTF-Annulated Hybrid Calix[n]pyrrole Sys-
tems 2663 Received: June 15, 2016
Published: October 18, 2016

© 2016 American Chemical Society 2641 DOI: 10.1021/acs.chemrev.6b00375


Chem. Rev. 2017, 117, 2641−2710
Chemical Reviews Review

2.6.3. Supramolecular Nonbonded Donor−Ac- on TTF that are attractive in the quest to make electronic
ceptor Ensembles of Porphyrin and TTFs 2682 devices,2−11 organic conductors,12 light-harvesting anten-
2.7. TTF-Functionalized Porphyrazines 2684 nas,13−15 optically active materials,16,17 and pigments.18
2.7.1. TTF-Annulated Unsymmetrical and Sym- To frame the seminal insights provided by the iconic R. S.
metrical Porphyrazines 2684 Mulliken and his charge-transfer theory and R. A. Marcus and his
2.7.2. Porphyrazines Peripherally Modified electron-transfer mechanism in the context of TTF chemistry, a
with TTFs through Saturated Spacers 2687 brief theoretical background is provided here.
2.7.3. Pyrazinoporphyrazine with Peripheral In an effort to create efficient donor−acceptor systems (e.g., a
TTF Units Tethered by Flexible Spacers 2688 TTF−TCNQ complex), the choice of the electron-donating and
2.8. TTF-Functionalized Phthalocyanines 2688 -accepting molecular units in the ensembles is critical.19 Relative
2.8.1. TTF-Annulated Symmetrical Phthalocya- to the acceptor (A), an electron-donor molecule (D) has, in
nines 2688 general, a lower ionization energy (Ip), thereby being oxidized
2.8.2. TTF-Crown-Ether-Functionalized Phtha- more readily. In contrast, an electron-accepting molecule (A) has
locyanines 2691 a larger electron affinity (Ea) and can be reduced at relatively low
2.8.3. TTF-Annulated Unsymmetrical Phthalo- potentials. Electronic coupling between the highest occupied
cyanines 2693 molecular orbital (HOMO) of the donor and the lowest
2.8.4. Phthalocyanines Containing Peripheral unoccupied molecular orbital (LUMO) of the acceptor
TTF Units Connected through Saturated molecules typically results in a degree of partial charge transfer ρ
Spacers 2694
diffuse ET diffuse
2.8.5. Norphthalocyanines Attached to TTFs D + A HoooooI [D···A] HooI [D ρ +···Aρ −] HoooooI D+ + A− (1)
through Saturated Spacers 2697
2.9. TTF-Annulated Subphthalocyanines 2700 Mulliken conceived that a diffusional encounter between an
3. Conclusions and Future Prospects 2701 electron donor, D, and an electron acceptor, A, would form a
Author Information 2702 reversible encounter complex (D···A) and worked to correlate
Corresponding Authors 2702 the extent of the interspecies interaction with the observed
Notes 2702 optical transition(s).20
Biographies 2702 In the case where the value of ρ is close to 1, a radical ion pair
Acknowledgments 2703 (D•+ and A•−) can be obtained through an electron-transfer
Abbreviations 2703 (ET) event. ET is a mechanistic description of the thermody-
References 2704 namic concept of a redox process, wherein the oxidation states of
both reaction partners change. The accompanying electron-
transfer state, [D•+ A•−] is often produced under the conditions
1. INTRODUCTION of reversible equilibrium. Therefore, electron back-transfer
Tetrathiafulvalene (TTF) is an effective π-electron-donating generally occurs as well. By far, the most accessible contribution
molecule that displays unique electrochemical behavior. These to the development of ET theory came from Marcus, whose
key attributes have resulted in its extensive use in the seminal contributions from 1956 onward were recognized by the
construction of organic conductors. A characteristic feature of Nobel Prize in Chemistry in 1992.21,22 Marcus’ electron-transfer
TTF, mirrored in a number of its derivatives, is that, upon formulation provided a major framework for understanding
stepwise oxidation, it gives rise to two sequentially oxidized electron-transfer dynamics.
species, namely, the TTF radical cation (TTF•+) and the TTF The charge transfer can be either thermally or optically
dication (TTF2+). The oxidation is thermodynamically reversible induced. It can be described by transitions between, or motions
(cf. Figure 1) within the accessible electrode potential range (viz., on, free energy surfaces (FESs).23 The free energy of the whole
electron donor−acceptor system, including the solvent environ-
ment, is given as a function of the reaction coordinate (Figure 2).
Following the Franck−Condon principle, electron motion is
much faster than nuclear motion. As a consequence, optically
induced charge transfer occurs from the minimum of the reactant
Figure 1. Reversible redox processes involving the neutral (TTF),
radical cation (TTF•+), and dication (TTF2+) species, respectively.
These redox changes endow TTF with unique photophysical and
electrochemical properties.

E1/21 = +0.34 V and E1/22 = +0.73 V vs Ag/AgCl in acetonitrile).1


TTF is also thermodynamically quite stable in a variety of
different chemical environments (except in the presence of
strong acids and oxidants). Thus, it is widely used as a building
block in the formation of supramolecular structures and organic
charge-transfer (CT) complexes. Functionalization of the TTF
core is often easily achieved and allows for synthetic integration
of the TTF unit into various D−A or D−π−A systems (where D Figure 2. Diabatic (dashed line) and adiabatic (solid line) free energy
stands for donor and A stands for acceptor). This has allowed the surfaces of an electron donor−acceptor system. The optically induced
creation of both inter- and intramolecular CT constructs based (green) and thermally induced (blue) charge transfers are shown.

2642 DOI: 10.1021/acs.chemrev.6b00375


Chem. Rev. 2017, 117, 2641−2710
Chemical Reviews Review

state generated by an increase in the energy along the vertical


direction to the product FESs, whereas thermally induced charge
transfer can occur only in the transition-state (TS) region. In the
case of thermally induced charge transfer, the system has to move
along the reaction coordinate to the TS region, resulting in an
activation barrier. The barrier height can be estimated for
optically induced charge transfer using geometric considerations.
The free energy of activation, ΔG‡ET, can be calculated from the
point where the two parabolas intersect
0
‡ (ΔG ET + λtotal)2
ΔG ET =
4λtotal (2)
The free energy difference between the minima of the initial
and final states is represented by ΔG0ET, and λtotal is the total
reorganization energy, which is composed of the inner-sphere
reorganization energies (λv) and outer-sphere (λo) reorganiza-
tion energies. Therefore, λv and λo represent the intramolecular
and solvent coordinate changes, respectively, in the optically
induced charge-transfer process, and the sum ΔG0ET + λtotal is the
total energy needed to induce optically driven charge transfer.
Whereas λv cannot be calculated in a simple manner, λo can
commonly be estimated according to the classic Marcus
expression, where the donors and acceptors are assumed to be
imbedded in a dielectric continuum

e2 ⎛ 1 1 ⎞⎛ 1 1 1⎞
λ0 = ⎜ − ⎟⎜ + − ⎟
4πε0 ⎝ n 2
D ⎠⎝ 2R+ 2R − d⎠ (3)
and the electron-donor and electron-acceptor redox centers R± Figure 3. (a) Schematic diagrams showing different regions described
are assumed to be spherical. The distance between these redox by the Marcus theory of electron transfer as applied to a D−A system.
centers is represented by d. The solvent continuum is described (b) Driving force dependence of log kET on different λ values.
by its refractive index n, permittivity D, and dielectric constant ε0.
In the adiabatic case, the free energy of activation, ΔG‡ET, is
reduced in comparison to the diabatic limit. This is a result of the Barrierless (i.e., optimal) electron transfer can occur if λtotal =
FES splitting in the TS region. ΔG0ET. In the normal region associated with Marcus theory, an
The rate constant of charge transfer, k, is governed by the increase in the free energy driving force, ΔG0ET, results in a
barrier height given by ΔGET ‡
. The probability P can be decrease in the activation energy, ΔG‡ET. A further increase in
rationalized by the time the system needs to cross the TS region ΔG0ET (≈ λtotal) yields the maximum rate of electron transfer.
versus the time the system needs to change from the reactant to Eventually, an increase in the driving force leads to an increase in
the product state. In the diabatic regime, the change of states is the activation barrier, ΔG‡ET, and the rate of electron transfer
slow as the result of motion required through the TS region. decreases. This portion of the FES is generally referred to as the
Thus, the probability for the transition from the reactant to the Marcus inverted region (Figure 3b).
product state is low. As a consequence, the overall charge transfer A number of electron-donor−acceptor systems containing
is limited by P, which, in turn, is proportional to the square of the porphyrin and phthalocyanine analogues have been prepared in
electronic coupling element V2. In the diabatic limit, the rate an effort to control electron-transfer processes. These frame-
constant of charge transfer, kna (na = nonadiabatic), can be works are attractive because of their inherent highly delocalized
described by the Arrhenius-type equation electronic structures and the relatively rigid π-scaffolds they
provide. For instance, porphyrins typically display relatively small
1 reorganization energies and undergo minimal structural and
k na = 4π 2ℏc 2
4π ℏcλtotalkBT solvation changes when used as components in ET couples.24−27
The electronic features of TTF and its derivatives also make
⎡ ℏc(λ 0 2⎤
total + ΔG ) them attractive for ET studies. Synthetic modifications can be
V 2 exp⎢ − ⎥
⎣ 4λkBT ⎦ (4) used to modulate their intrinsic donor−acceptor characteristics,
making TTF-based systems attractive as strong electron donors
where kB and ℏ are the Boltzmann and Plank constants, in both static and dynamic supramolecular motifs. Pyrrolo-TTF
respectively; c is the speed of light in a vacuum; T is the systems have had a particularly important role to play in the latter
temperature; and ΔG0 is the free energy difference between the context.
reactant and product states. In this case, the logarithm of the ET In recent years, considerable effort has been devoted to the
rate constant (log kna) is related parabolically to the ET driving construction of various supramolecular architectures displaying
force (negative ET free energy change) between the electron controlled molecular movement triggered by external stimuli
donor and acceptor and the ET reorganization energy, λtotal, that such as specific acids, anions, and a variety of guest molecules.
is, the energy required to structurally reorganize the donor, For the most part, the designs have been based on the use of
acceptor, and their solvation spheres upon ET (Figure 3a). receptors with strong affinity toward guest molecules and have
2643 DOI: 10.1021/acs.chemrev.6b00375
Chem. Rev. 2017, 117, 2641−2710
Chemical Reviews Review

Chart 1. Representative TTF-Annulated Oligopyrrolic Macrocycles

relied on a variety of noncovalent interactions (e.g., hydrogen- highly conjugated macrocyclic core (e.g., porphyrin, porphyr-
bonding, electrostatics, and π−π D−A-based effects). TTF has azine, and phthalocyanine) allows TTF chemistry to be
proved to be a near-ideal molecular component in such systems. combined with porphyrin chemistry to produce hybrid systems
It has been exploited in the design of many different with many interesting photophysical and electrochemical
supramolecular architectures, including molecular ma- properties. Therefore, these systems are a primary point of
chines,28−32 molecular flasks,33,34 molecular tweezers,35−37 and focus in this review.
molecular switches.38−46 TTF-derived compounds have been widely described in book
Some of the resulting constructs have potential applications as chapters,12,47 accounts,48 feature articles,27,49,50 and review
sensors for organic neutral guests and hazardous nitroaromatic articles.1,51−72 A few of the recent review articles60,73 discuss
compounds (vide infra). Generally, the observed sensing TTF oligomers and fused TTF derivatives. Some of the
behavior arises from the formation of supramolecular ensembles reviews74−79 emphasize the synthesis and properties of various
by specifically designed guests with TTF-annulated hosts, which TTF derivatives with different functionalities. There are also a
can be detected spectroscopically or electrochemically at the couple of review articles on TTF-based cyclophanes and cage-
micro- to nanomolar levels. These hybrid systems are particularly motif systems,80 as well as those detailing D−A assemblies and
important because they contain both a redox-active site and a their potential use in the areas of molecular materials and
chromophoric unit that jointly or separately provide for an devices.81−85 Other reviews focus on the charge-transfer and
efficient spectroelectrochemical or photophysical response to the energy-transfer properties of TTF systems and their potential in
targeted analyte. In many instances, TTF functionalization of a the areas of organic electronic materials86−89 and magnetic
receptor serves to enhance the overall electron-donor strength of materials.90 However, to the best of our knowledge, a detailed
the resulting supramolecular construct compared to non-TTF- review focused on TTF-annulated oligopyrrolic macrocyclic
based systems. This modification often provides for enhanced systems has yet to appear in the literature. Given the utility of
hydrogen-bonding or π−π interactions with anions and electron- these systems in the areas of ET study, molecular recognition,
deficient guests. Moreover, as detailed further below, fusion of a supramolecular materials chemistry, and sensor development, we
TTF subunit to a calix[4]pyrrole core can give rise to molecular believe that a comprehensive review is warranted. Therefore, our
containers with concave openings suitable for encapsulating vision is to highlight the most recent developments involving
spherical guest molecules. In addition, TTF annulation to a various TTF-annulated oligopyrrolic macrocyclic compounds,
2644 DOI: 10.1021/acs.chemrev.6b00375
Chem. Rev. 2017, 117, 2641−2710
Chemical Reviews Review

Scheme 1. Synthetic Routes Leading to Various TTF-Annulated Pyrroles

including TTF-calix[n]pyrroles (where n = 4, 5, and 6), TTF- macrocycles as reported over the past two decades. We also
porphyrins, so-called exTTF-porphyrins, TTF-porphyrazines, summarize work focused on TTF-fused porphyrins, as well as
TTF-phthalocyanines, and TTF-subphthalocyanines, as well as several extended TTF analogues that contain formal porphyrin
several closely related structures (Chart 1). bridges (termed exTTF porphyrins).
In particular, we have made efforts to discuss in detail the All of these oligopyrrolic macrocyclic compounds containing
supramolecular recognition and charge-transfer properties of TTF(s) are potential electron donors capable of complexing a
nonaromatic TTF-pyrrole systems, with an emphasis being wide range of electron-poor guests. Their sensing behavior
placed on the detection and sensing behavior of these mainly reflects their ability to form supramolecular D−A-type
2645 DOI: 10.1021/acs.chemrev.6b00375
Chem. Rev. 2017, 117, 2641−2710
Chemical Reviews Review

Chart 2. Chemical Structures of Various TTF-Annulated Calix[4]pyrrole Receptors and the Reference Compound C[4]P

assemblies. As a general rule, the efficiency whereby such D−A The resulting TTF-annulated pyrrole derivative represents a
supramolecular ensembles are formed depends strongly on the useful new building block that can bring both rigidity and
electronic nature of the components, the three-dimensional electron-donating character to various types of oligopyrrolic
arrangement of the planar TTF units within the macrocyclic macrocycles (e.g., porphyrins, calixpyrroles). When it is
receptor and guest, and the environment surrounding the self- incorporated into a calix[n]pyrrole (n = 4, 5, and 6) framework,
assembled complex. higher binding affinities toward anions as the result of NH−
In the first part of this review, we discuss the design and anion hydrogen-bonding interactions are generally observed.
synthesis of TTF-annulated pyrrolic precursors before turning to The peripheral TTF unit can also be utilized as secondary
calix[4]pyrroles containing TTF subunits, as well as their higher interaction sites by providing a set of electron-rich π-surfaces.
analogues. TTF-calix[n]pyrroles are structurally flexible systems From the viewpoint of the pyrrolic TTF building block and
wherein a number of TTF units are fused to the periphery of the systems created from it, modification of the terminal side of the
core macrocycle. As detailed further below, they are often capable TTF unit (i.e., that farthest from the pyrrole NH proton) can
of sensing anions, detecting nitroaromatic explosives, and allow for further fine-tuning of the electronic structure. Common
recognizing a variety of guest molecules. The electrochemical functionalized TTF pyrrole derivatives are provided by benzo
response and photophysical behavior of these systems are also annulation and thioalkyl substitution. These two substitution
discussed thoroughly. In some cases, the reversible switching of patterns have received particular attention because they provide a
an electron-transfer process within a supramolecular ensemble as reasonable balance between solubility in organic solvents and
a function of guest molecules is considered. The effects of chemical stability of the derivatives. However, other substituted
metalation and various guest inputs and their effects on the analogues are known.
structure and properties of a few highly flexible hybrid A key step in making a TTF-annulated oligopyrrolic
calixpyrrole systems are also highlighted. In the second part, macrocyclic system is the synthesis of a TTF-fused pyrrole
we focus on TTF-fused porphyrins, as well as higher-order TTF- precursor specifically designed for the particular macrocyclic
derived porphyrins and their artificial photosynthetic applica- target in mind. Approaches used to prepare a few well-studied
tions. We also discuss the use of specific guest inputs to control TTF-annulated pyrroles (e.g., PnS-TTF-pyrrole,91 MeS-TTF-
charge-transfer (CT) phenomena in the case of TTF-fused pyrrole,91,92 PrS-TTF-pyrrole,93,94 BTTF-pyrrole,95 TTTF-
porphyrins. The interesting ground-state and excited-state pyrrole,95 TTF-dipyrrole91,92,94) are outlined in Scheme 1.
dynamics, as well as electrochemical behavior, that originate Typically, the synthesis of TTF-pyrroles involves a CC
from the presence of a chromophoric porphyrin subunit and one bond-forming reaction carried out in triethylphosphite, P(OEt)3,
or more redox-active TTF groups are also detailed. The involving the coupling between an appropriately chosen 1,3-
electrochemical and optical properties of various TTF-fused dithiole-2-thione and N-tosyl-(1,3)-dithiolo[4,5-c]pyrrol-2-one
porphyrazines, phthalocyanines, and subphthalocyanines are (OX) derivatives. This is an approach that has proved useful in
discussed thoroughly in the last part of this review. Every effort obtaining unsymmetrical cross-coupling products (e.g., PrS-
has been made to cover the literature through early 2016. TTF-pyrrole, BTTF-pyrrole, and TTTF-pyrrole). The requisite
heterocoupling often proceeds well because the corresponding
2. TTF-ANNULATED OLIGOPYRROLIC MACROCYCLIC symmetrical products require a homocoupling of either the 1,3-
COMPOUNDS dithiole-2-thione or OX components. Although these latter
homocouplings can be exploited to obtain species, such as TTF-
2.1. TTF-Annulated Calix[4]pyrroles bridged bis-pyrroles (TTF-dipyrrole), to date, the products in
Annulation of a TTF unit onto a pyrrole scaffold through the question have not been exploited extensively in the preparation
β,β′-positions has a significant effect of the acidity of the pyrrolic of more elaborated products. This stands in contrast to what is
NH proton because of the inductive effect of the sp2-hybridized S true for the pyrrolic heterocoupling products. Generally, the
atoms and the intrinsic electron-rich nature of the TTF subunit. pyrrole nitrogen is tosylated. Detosylation using sodium
2646 DOI: 10.1021/acs.chemrev.6b00375
Chem. Rev. 2017, 117, 2641−2710
Chemical Reviews Review

Scheme 2. Synthesis of Various TTF-Calix[4]pyrroles

methoxide (NaOMe) proceeds well, with the result that several tetramerization of PrS-TTF-pyrrole in acetone; after purification,
key TTF-pyrrole precursors are now routinely synthesized in the this product was isolated in 18% yield.
authors’ laboratories on gram scales. The anion-recognition properties of receptor 1 were
2.1.1. TTF-Annulated Calix[4]pyrroles for Anion Sens- investigated by means of NMR spectroscopic titrations and
ing. In 2003, Becher and co-workers reported the first electrochemical analyses. The Job’s plots obtained upon the
calix[4]pyrrole system fused with TTF, the mono-PnS-TTF- titration of 1 with a number of anion salts exhibited maxima at a
calix[4]pyrrole receptor 1 (Chart 2), which was produced as a mole fraction of 0.5, as would be expected for the formation of
possible electrochemical sensor for anions.96 Compound 1 was 1:1 complexes. An electrochemical response toward different
produced through a straightforward synthetic procedure that anions was seen for 1 and was considered to be enhanced by the
involved the BF3·OEt2-catalyzed macrocyclization of a bis- presence of a relatively more acidic −NH proton present in the
hydroxymethyltripyrrane with pentylthio-TTF-pyrrole (PnS- PnS-TTF-pyrrole subunit because of the presence of a TTF
TTF-pyrrole) in a 1:1 stoichiometric ratio, as outlined in fragment that is electron-rich relative to a normal pyrrole
Scheme 2A. (C[4]P; Chart 2). The binding constants corresponding to the
Representative target-oriented synthetic routes used to interaction of 1 with halide anions F−, Cl−, and Br− (as the tetra-
prepare other TTF-annulated calix[4]pyrrole receptors are n-butylammonium, TBA+, salts) were determined from 1H NMR
shown in Scheme 2B. For instance, the symmetrical TTF-fused spectroscopic titration data obtained at 300 K; the results are
calix[4]pyrrole 4 was synthesized by the acid-catalyzed cyclo- summarized in Table 1. The underlying experiments involved
2647 DOI: 10.1021/acs.chemrev.6b00375
Chem. Rev. 2017, 117, 2641−2710
Chemical Reviews Review

Table 1. Binding Constant Values Corresponding to the result of the strong hydrogen-bonding interactions between the
Interaction of Receptor 1 with Various Anions pyrrolic protons and the bound anionic guest (Figure 5).
anion Kab (M−1)

F 2.1 × 106
Cl− 1.2 × 105
Br− 7.6 × 103
a
Underlying NMR spectroscopic titrations were carried out using
tetra-n-butylammonium (TBA+) salts of the corresponding halides at
300 K. Binding constant values were obtained by the nonlinear curve-
fitting computer program EQNMR and is the average of two
independent experiments using the chemical shifts of the three
−NH protons as the probe signals. The estimated error was <10%.
b
CD3CN/0.5% v/v D2O.

monitoring the change in the chemical shifts of the resonances


associated with the three nonequivalent −NH protons of the
macrocyclic core as the amount of the corresponding
tetrabutylammonium halide salt was increased. The data were
then fitted to a 1:1 binding profile.
It was also demonstrated that the redox-active TTF unit of this
receptor 1 respond in diagnostic fashion upon complexation with
anions such as Br− in MeCN. This is evident (Figure 4) in terms

Figure 5. Two conformational structures of tetrakis-TTF-calix[4]-


pyrrole, 4. The conversion between these limiting forms is thought to
underlie much of their interesting chemical behavior. (This figure is
based on one that originally appeared in ref 98 but has been redrawn.)

The distinct dynamic motion of TTF-calix[4]pyrrole demon-


Figure 4. (a) Cyclic voltammograms of receptor 1 (0.5 mM) recorded
strated in Figure 5 shows that, in the absence of anions, the
upon the incremental addition of tetra-n-butylammonium bromide system adopts a 1,3-alternate conformation, whereas in the
(TBABr) in MeCN at 298 K using tetra-n-butylammonium presence of anions (e.g., Cl−), a cone-shaped conformer is
hexafluorophosphate (TBAPF6, 0.4 M) as the supporting electrolyte thermodynamically preferred as a result of strong hydrogen-
at a scan rate of 0.5 V s−1. (b) CV titration results showing the cathodic bonding interactions with the bound Cl− guest.
shift of the first oxidation potential of receptor 1 seen upon the gradual In the case of compounds 2 and 4, X-ray crystallographic
addition of Br− ions. Note that the cathodic shift of the first oxidation analyses (Figure 6) served to reveal that, in the absence of any
potential is independent of the second oxidation process. (Reproduced added anion, the systems adopt a 1,3-alternate conformation
with permission from ref 96. Copyright 2003 Wiley-VCH Verlag similar to those of other normal nonannulated calix[4]pyrrole
GmbH.) systems, such as C[4]P.
Anion-binding studies carried out in organic media [1,2-
dichloroethane (DCE)] using soluble anion salts served to
of a cathodic displacement of the first oxidation potential (E1/21), demonstrate that incorporation of one or more electron-rich
along with an increase in the current intensity associated with the
second TTF-based oxidation potential (E1/22), as determined by
cyclic voltammetry (CV). This makes receptor 1 a very efficient
redox-responsive chemosensor for halide anions in this organic
medium.
Jeppesen and co-workers reported a series of TTF-substituted
calix[4]pyrrole receptors, 2−4, which were also tested as
potential electrochemical sensors.97 Among these, receptor 4
exhibited many interesting chemical and photophysical features
that were ascribed in large measure to its ability to adopt two
different limiting conformations, namely, the (i) 1,3-alternate
and (ii) cone conformers in the absence and presence of anions,
respectively. In the absence of anions, the 1,3-alternate form is Figure 6. Solid-state structures of (a) mono-TTF-calix[4]pyrrole (2)
energetically favorable, as it is for most other calix[4]pyrroles, and (b) tetrakis-TTF-calix[4]pyrrole (4), showing the 1,3-alternate
whereas in the presence of coordinating anions, such as Cl−, conformers. (This figure was redrawn using data that were originally
these systems adopt the corresponding cone conformation as a published in ref 97.)

2648 DOI: 10.1021/acs.chemrev.6b00375


Chem. Rev. 2017, 117, 2641−2710
Chemical Reviews Review

TTF subunits into the calix[4]pyrrole backbone improves the


anion-binding abilities of the receptors while also affecting their
selectivity (Table 2).

Table 2. Binding Constants (Ka, M−1) Corresponding to the


Interactions between Receptors 2, 3, and 4 and Different
Anions as Determined by Isothermal Titration Calorimetry
(ITC) at 298 K in DCE
anion 2 3 4
Cl− 1.2 × 105 6.6 × 105 2.5 × 106
Br− 2.2 × 103 8.3 × 103 5.8 × 104
CN− 8.5 × 104 2.6 × 105 1.1 × 106
NO2− 6.9 × 104 2.8 × 105 1.2 × 106
MeCO2− 6.5 × 104 3.9 × 105 1.3 × 106
a
Estimated errors are <10%. bReceptors 2, 3, and 4 were titrated with
the tetra-n-butylammonium salts of the anion in question so as to
obtain the heat effects corresponding to complexation. The net heat
effect was determined by subtracting the heat traces for the
appropriate background titration.

For instance, the binding constant for Cl− (as the TBA+ salt) in
the case of 4 is significantly enhanced (i.e., Ka = 2.5 × 106 M−1)
compared to that of either C[4]P or 2 (i.e., 3.5 × 104 M−1 for
C[4]P and 1.2 × 105 M−1 for 2). As the number of TTF-pyrrole Figure 7. Cyclic voltammograms of 0.25 mM 1,2-dichloroethane
units increases, the binding constants for anions likewise (DCE) solutions of receptors (a) 2, (b) 3, and (c) 4 recorded vs Fc+/Fc0
increase. These findings are consistent with the notion that the in the presence of increasing quantities of TBAX (in the case of
presence of electron-donating TTF units within this class of receptors 2 and 4, X = Cl−; for receptor 3, X = Br−). Experiments were
receptors serves to strengthen the NH−anion hydrogen-bonding carried out at 298 K using TBAPF6 (0.1 M) as the supporting electrolyte.
interactions. The effect can be further enhanced by the presence (Reproduced with permission from ref 97. Copyright 2006 American
of electronegative sp2-hybridized sulfur atoms in this particular Chemical Society.)
set of receptors. This putative electronic effect was also seen in
the case of other anions (e.g., Br−, CN−, NO2−, and MeCO2−). receptor 4, a shift in the first oxidation potential to a limit of ΔE =
Of note is that halide complexation could also be monitored by −30 mV was seen (Figure 7c) upon the incremental addition of
electrochemical means. For instance, clear potential shifts in the n-Bu4NCl. The largest shift in the first oxidation potential, ΔE =
TTF-based redox waves were observed as a function of anion −145 mV, was observed in the case of receptor 2 when titrated
concentration (Figure 7). with Cl−. Although the degree of response clearly varies in
It was observed that, upon addition of 0.2 equiv of n-Bu4NCl qualitative terms, all of these TTF-functionalized macrocyclic
(as the source of Cl−) to receptor 2 in 1,2-dichloroethane at 298 receptors behave as rudimentary electrochemical sensors for the
K (Figure 7a), a new peak at E1/21 = −0.11 V appeared in the Cl− and Br− anions.
region associated with the first oxidative event. This new peak is 2.1.2. TTF-Annulated Calix[4]pyrroles for Sensing
ascribed to oxidation of the complex formed as the result of Neutral Aromatic Electron-Deficient Guests. The rationally
chloride-anion binding, namely, 2·Cl−. Further addition of Cl− designed TTF-annulated calix[4]pyrrole system 4 also acts as an
ions led to a monotonic decrease in the intensity of the original effective receptor for 1,3,5-trinitrobenzene (TNB), tetrafluoro-p-
peak at E1/21 = +0.032 V assigned to oxidation of the benzoquinone, tetrachloro-p-benzoquinone, and p-benzoqui-
uncomplexed form (i.e., free receptor 2) with a concurrent none in CH2Cl2 solution.98 This neutral substrate recognition
increase in the intensity of the new peak appearing at E1/21 = process can be blocked by the addition of chloride anion.
−0.11 V. A total displacement (ΔE) of −145 mV of the first Receptor system 4 was designed to take advantage of the fact that
oxidation potential was observed in this case. The buildup in the calix[4]pyrroles generally exist in the 1,3-alternate conformation
signal at E1/21 = −0.11 V was considered to reflect an increase in in the absence of anions. In this latter conformation, each pair of
the relative concentration of the complex (2·Cl−) at the expense identical TTF electron donors is expected to hold an electron-
of that of the uncomplexed receptor 2. In the case of receptor 3, deficient guest in a sandwich-like fashion through D−A−D-
gradual addition of n-Bu4NBr (as the source of Br− anion) under based CT interactions (Figure 8a). Additional stabilization is also
identical experimental conditions resulted in a slight negative provided by pyrrole NH−substrate hydrogen-bonding inter-
shift (Figure 7b) of the first oxidation potential (E1/21) associated actions. Based on a range of experimental findings, including X-
with the TTF units. It was also noticed that, upon increasing the ray structural analyses, limiting binding modes were deduced for
Br− concentration, the intensity of current associated with the receptor 4 interacting with various electron-deficient species,
second oxidation wave gradually increased. This was rationaized including the model electron-deficient nitroaromatic explosive
in terms of the Br− ions being oxidized to Br2 under these 1,3,5-trinitrobenzene (TNB) and tetrafluoro-p-benzoquinone.
conditions. The displacement of the first oxidation potential These interactions are shown for TNB in Scheme 3, along with
reached a limit of ΔE = −70 mV at the point where the changes seen upon changes in treatment conditions. An X-ray
approximately 1 stoichiometric equiv of Br− ion had been structure of the complex between 4 and 2 equiv of TNB is shown
added to the solution containing receptor 3. In the case of in Figure 8a.
2649 DOI: 10.1021/acs.chemrev.6b00375
Chem. Rev. 2017, 117, 2641−2710
Chemical Reviews Review

Positive Homotropic Allosteric Binding Properties of TTF-


Calix[4]pyrrole Derivatives with Electron-Deficient Nitro-
aromatic Guests. The design and construction of artificial
receptors displaying nonlinear amplification as a result of an
initial guest binding event represents a well-appreciated
challenge in supramolecular chemistry. As above, compound 4
was found to bind polynitroaromatic explosives with a 1:2
(receptor/guest) stoichiometry in organic solvents in the
absence of a coordinating anion, such as Cl−. However, it was
also found that the binding affinity of 4 for electron-deficient
guests is somewhat modest. As a consequence, relatively high
guest concentrations are needed to produce a naked-eye-
detectable colorimetric response. To overcome this deficiency,
aromatic (thiophene and benzene) annulated tetrakis-TTF-
calix[4]pyrroles, namely, 5 and 6, were synthesized, and their
abilities to bind polynitroaromatic explosives, including TNB,
picric acid (TNP), and 2,4,6-trinitrotoluene (TNT), were tested.
A careful analysis of the colorimetric changes observed when
TNB, TNP, and TNT were allowed to react with 4−6 revealed a
positive homotropic allosteric effect (K1 < K2).95 Figure 9
illustrates the proposed origin of this positive homotropic
allosteric effect in the case of receptor 5 and TNP. Briefly, the
effect is explained by the fact that binding of a first guest forces
the inherently flexible host systems to adopt a more rigid 1,3-
Figure 8. (a) X-ray structure of the supramolecular complex (TNB)2⊂4. alternate structure. This, in turn, leads to a loss of rotational
(b) UV−vis−NIR spectral changes seen for mixtures of 4 and TNB freedom, resulting in a preorganized receptor structure that is
under various experimental conditions. (Reproduced with permission better able to encapsulate the second target guest molecule.
from ref 98. Copyright 2004 American Chemical Society.) Based on quantitative analyses of visible spectroscopic titration
results, receptor 6 was found to display a greater degree of
positive homotropic allostery than the other systems, for which a
general trend of 6 > 5 > 4 was found (Table 3).
An intermolecular CT band at λmax = 677 nm was seen in the Zhu et al. reported a TTF-calix[4]pyrrole 6 that displayed a
UV−vis−NIR absorption spectrum (Figure 8b) when 4 was colorimetric response toward polynitroaromatic explosive both
treated with TNB. This band disappeared upon addition of Cl−. in chloroform solution and when incorporated on a polymer
This was taken as evidence that an anion-induced conformational microcantilever with an integrated deflection-sensing element.99
change leads to a guest-release process and a breakup of the initial This work represents an approach to creating a practical device
CT complex. After removal of the anion by means of an aqueous capable of stand-off detection of nitroaromatic explosive vapors.
wash, the initial sandwich structure was restored, along with the It was also expected that the microcantilever systems would
low-energy CT-visible absorption feature. display enhanced sensitivity and improved detection limits

Scheme 3. Proposed Supramolecular Host−Guest Chemistry Involving Receptor 4 and TNB Guest Observed under Different
Experimental Conditions

2650 DOI: 10.1021/acs.chemrev.6b00375


Chem. Rev. 2017, 117, 2641−2710
Chemical Reviews Review

Figure 9. Proposed origin of the positive homotropic allosteric effect observed in the case of 5 and TNP. (This figure was redrawn using data that were
originally published in ref 95.)

Table 3. Microscopic Association Constants, and achieved. This represents an improvement in the detection limit
Cooperativity Parameters for the Receptors 4, 5, and 6 with of at least 30-fold compared to the colorimetric detection limit of
Various Nitroaromatics (TNB, TNP, and TNT) 0.3 ppm for TNB measured in chloroform solution using
compound Ka (M−2) n K1 (M−1) K2 (M−1) K2/K1 receptor 6. This study provides an indication that stand-alone
4·2TNB 4.3 × 103
1.27 3.9 × 102
1.4 × 103
3.6 colorimetric receptors could potentially be translated into
4·2TNP 3.8 × 103 1.30 2.8 × 102 1.2 × 103 4.1 practical devices with improved sensitivities by incorporating
4·2TNT 3.3 × 102 1.23 5.9 × 101 2.0 × 102 3.3 them into cantilever sensor arrays.
5·2TNB 1.5 × 105 1.34 2.8 × 103 1.7 × 104 6.2 Self-Assembly Driven Sensing of TNB with a TTF-Calix[4]-
5·2TNP 9.1 × 104 1.34 1.7 × 103 1.1 × 104 6.5 pyrrole Receptor. Nielsen and Stein reported the synthesis and
5·2TNT 1.2 × 104 1.31 5.7 × 102 2.6 × 103 4.5 binding properties of a TTF-calix[4]pyrrole receptor 7 appended
6·2TNB 3.4 × 106 1.70 1.3 × 103 3.1 × 104 24 with an electron-deficient 3,5-dinitrobenzoate guest moiety.100
6·2TNP 3.7 × 106 1.86 6.4 × 102 2.0 × 104 31 Preliminary spectroscopic studies revealed that the receptor self-
6·2TNT 2.3 × 104 1.45 3.2 × 102 2.8 × 103 10
a
associates into a dimer (7·7) at high concentration (Scheme 4).
Based on absorption spectroscopic titrations of the respective hosts The self-association of the receptor leads to preorganization in its
(0.20 mM) in CHCl3 at 298 K. bEstimated error for calculated binding
constants is ≤12%. 1,3-alternate conformation. As a result, dimer 7·7 displays an
approximately 2-fold higher binding affinity for nitroaromatic
relative to what was previously been found in the solution state. analytes (e.g., TNB) than the model receptor 4. This binding, in
In fact, a detection limit for TNB vapor of less than 10 ppb was turn, leads to the formation of a supramolecular ensemble

Scheme 4. Proposed Mechanism for the Self-Assembly of a TTF-Calix[4]pyrrole Derivative Driven by Complexation

2651 DOI: 10.1021/acs.chemrev.6b00375


Chem. Rev. 2017, 117, 2641−2710
Chemical Reviews Review

TNB2⊂7·7 that is stabilized through CT and hydrogen-bonding switch” consisting of the asymmetric TTF-calix[4]pyrrole
interactions. derivative 8 that displayed responsive binding behavior in the
UV−vis−NIR spectral titrations with various anions (as their case of TNB recognition.102 This receptor consists of three
TBA + cation salts) were carried out at two different identical TTF-pyrrole units and a fourth TTF-pyrrole unit
concentrations of receptor 7.101 It was found that, at higher appended with a phenol moiety. It is worth noting that the
concentrations [(0.15−0.40) × 10−3 M], the proposed anion- asymmetric system 8 was prepared by simply condensing two
binding events could be monitored spectroscopically in terms of different TTF-annulated pyrroles in accord with the normal one-
the progressive disappearance of the CT band centered at 560 pot synthetic strategy used to produce analogue 4. The host−
nm that occurs upon the incremental addition of coordinating guest interactions between 8 and TNB were studied by means of
anions, such as chloride. At lower concentrations of 7 [(6−7) × absorption and 1H NMR spectroscopies. The underlying studies
10−6 M], the absorption band centered at 326 nm was used to revealed that the molecular receptor could be switched between
determine the binding constants. The resulting values are given limiting locked and unlocked states by using base and acid as the
in Table 4. An inspection of the data in this table reveals that the inputs (Scheme 5).
In the unlocked state, the receptor is able to accommodate two
Table 4. Binding Constants (Ka, M−1) Corresponding to the TNB guest molecules through strong hydrogen-bonding
Interactions between the Receptor System 7 and the Parent interactions with the NH protons, as well as interactions
System 4 with Different Anions as Determined by Absorption between the electron-donating TTF units and the electron-
Spectroscopy and Isothermal Titration Calorimetry (ITC) in deficient TNB guest molecules. In contrast, TNB guests are not
CH2Cl2 at 298 K able to bind appreciably to the receptor in the locked state in
anion 7 (326 nm)d 7 (600 nm)d 4 (ITC) which the phenolate anion strongly coordinate with the NH

Cl 2.55 × 106
8.5 × 105
2.5 × 106 protons to construct a partial cone-like conformer. The addition
Br− 5.55 × 104 3.5 × 104 5.8 × 104 of acid serves to protonate the phenolate anion and restores TNB
CN− 4.65 × 105 2.4 × 105 1.1 × 106 binding.
MeCO2− 5.35 × 106 nde 1.3 × 106 Coordination-Driven Switching of a Preorganized and
a
Estimated errors are <15%. bReceptors 7 and 4 were titrated with the Cooperative TTF-Calix[4]pyrrole Receptor. Nielsen and co-
tetrabutylammonium salts of the anion so as to obtain the heat effects workers also reported a metal-ion-coordination-triggered switch-
corresponding to complexation. cNet heat effect was determined by ing system103 that consists of a preorganized asymmetric TTF-
subtracting the heat traces for the appropriate background titration. calix[4]pyrrole receptor, 9, bearing an appended pyridine moiety
d
Experiments were monitored at different wavelengths. eNot (Scheme 6). This system utilizes intramolecular hydrogen
determined. bonding between the pyrrole NH protons and the pyridine
nitrogen atom to switch off TNB binding in the 1,3-alternate
association constants obtained from the absorption spectro- conformation and metal-based complexation of the pyridine
scopic studies carried out at low receptor concentrations match moiety to turn it back on.
well those obtained from isothermal titration calorimetry (ITC) Absorption and 1H NMR spectroscopic data analyses along
studies of the parent system 4. In contrast, the binding constants with theoretical calculations revealed that the system has the
obtained at high concentrations of 7 are lower. This is taken as ability to complex electron-deficient nitroaromatics, such as
evidence that, at high concentrations, self-association serves to TNB, and that the binding events could be reversibly modulated
produce dimer 7·7, thus diminishing the inherent propensity to by using Zn2+ as an external stimulus. In the free state (i.e., in the
recognize anions. absence of Zn2+), the receptor system is locked into a relatively
Acid−Base-Induced Switching of a TTF-Calix[4]pyrrole higher level of preorganization as a result of the intramolecular
Receptor and Evidence for a Positive Cooperative Effect. hydrogen-bonding interactions (self-complexation) between the
Nielsen et al. also reported an acid−base-controllable “molecular pyrrole NH protons of the TTF-calix[4]pyrrole skeleton and the

Scheme 5. Acid−Base-Controlled Supramolecular Guest Binding Using an Unsymmetrical Receptor

2652 DOI: 10.1021/acs.chemrev.6b00375


Chem. Rev. 2017, 117, 2641−2710
Chemical Reviews Review

Scheme 6. Proposed Mechanism for the Coordination-Driven Host−Guest Complexation Seen in the Case of Receptor 9

pyridine moiety attached to one of the TTF side arms. In this Scheme 7. Illustration of Proposed Binding Mode of DNT
particular conformation, the receptor displays negative cooper- Guests Seen in a Cantilever System Based on Receptor 4
ativity in that initial TNB binding is more facile than subsequent
binding events. Specifically, the binding of a second molar
equivalent of TNB requires displacement of the appended
pyridine, which is held in place by hydrogen-bonding
interactions as noted above. Addition of hexafluoroacetylaceto-
nato Zn(II), [Zn(hfac)2], to the system provides a Lewis acid
(Zn2+) that interacts with the appended pyridine moiety. This
allows the parent receptor to revert to the more conformationally
flexible state characteristic of calix[4]pyrroles. Therefore,
addition of Zn2+ induces positive cooperativity in the binding
of TNB by receptor 9.
Bosco et al. reported a modified DVD platform for high-
throughout detection of 2,4-dinitrotoluene (DNT) wherein
silicon cantilevers were functionalized with the receptor system
4.104 Scheme 7 shows the proposed guest binding mode for this
cantilever system.
It was observed that a significant cantilever deflection occurred
when the functionalized cantilever was exposed to DNT.
Moreover, a change in the resonance frequency was detected,
presumably reflecting the added mass resulting from the
encapsulation of DNT within the cantilever system. The DNT
detection capability of this functionalized cantilever system,
which was standardized using reference cantilever systems, was
verified both in aqueous media and in the gaseous phase. Very
good reproducibility was seen, with over 80% of the independent
cantilevers displaying the same bending behavior when
monitored under the same conditions but using different
experimental batches (Figure 10).
2.1.3. TTF-Annulated Calix[4]pyrroles for Sensing
Spherical Guests. Jeppesen and co-workers developed a self- Originally, it was proposed that the fullerene binding was in
assembled multicomponent molecular host system based on 4105 the form of a 2:1 receptor/substrate complex (Scheme 8). This
that was found to interact with the electron-deficient acceptor inference was based on continuous-variation studies (Job plots)
C60 in the presence of a chloride anion source. wherein the total concentration of 4·Cl− and C60 was kept
2653 DOI: 10.1021/acs.chemrev.6b00375
Chem. Rev. 2017, 117, 2641−2710
Chemical Reviews Review

dichloromethane produces a brown solution. This solution is


characterized by the presence of a weak, broad absorption band
centered at about λmax = 748 nm (ε = 500 M−1 cm−1) in the
visible spectral region (Figure 11b, curve iv), as would be
expected if only minimal intermolecular interactions between
C60 and 4 existed. However, the addition of 5 equiv of TBACl to
this 2:1 mixture of 4 and C60 resulted in an instantaneous color
change from brown to green and the appearance of a strong CT
absorption band (Figure 11b, curve iii, centered at λmax = 725 nm;
ε = 1900 M−1 cm−1). Such a finding provides support for the
proposal that C60 is encapsulated only after 4 undergoes a change
from the 1,3-alternate conformation to the corresponding
chloride-bound cone conformation (i.e., 4·Cl−).
Electrochemical investigations revealed that, upon complex
formation, the first reduction of C60 shifts by −74 mV relative to
that of native C60. This finding supports the proposed donor−
acceptor interactions and the partial shifting of electron density
from the electron-rich TTF subunits to the electron-deficient C60
guest (Table 5). Furthermore, time-resolved femtosecond
transient absorption spectroscopy revealed that, upon photo-
Figure 10. (a) Gas-phase detection of DNT. Each curve reports the excitation, the electron-transfer rate between 4 and C60 is high
average differential signal between a set of 4-treated cantilevers (16 (i.e., the half-life for ET on the order of 1012 s−1).
cantilevers) and a set of reference cantilevers (8 cantilevers). Toluene The formation of the proposed supramolecular complex can
was used as a control analyte. (b) Solution-phase detection of DNT be easily inferred from the prominent visual color changes
guests studied using a continuous flow method in aqueous media. The relative to its constituents. No significant absorption was noticed
plots were generated by averaging the signals obtained using one for the receptor system 4 and the C60 guest molecule beyond λmax
functionalized chip (8 cantilevers) and one blank reference chip (8 ≥ 550 nm. However, when 0.5 equiv of C60 was added to receptor
cantilevers). The vertical gray line indicates the point where the system 4 in the presence of a Cl− anion source (which served to
is exposed to an aqueous DNT solution. (Reproduced with permission transform the 1,3-alternate structure to the corresponding bowl-
from ref 104. Copyright 2012 Elsevier Ltd.)
shaped conformer), the color of the solution turned green
immediately on the laboratory time scale as a result of the
constant. However, it was subsequently appreciated106 that, formation of a more ordered multicomponent supramolecular
because the concentrations of both 4 and Cl− were varied during ensemble that is now known to have a net stoichiometry of
the underlying titrations, a systematic error was introduced that C60⊂(4·Cl−) [and not C60⊂(4·Cl−)2 as originally assigned]. The
led to the assignment of an incorrect binding stoichiometry. This absorption spectrum of this complex gives a strong CT band with
followup study provides a discussion of some of the caveats λmax = 725 nm (Figure 11b). Addition of 2 equiv of the yellow
associated with using Job plots to infer stoichiometries when receptor 4 in its 1,3-alternate conformation (that is, in the
more than two guests can be bound to a single receptor. A absence of chloride anions) to a magenta solution of 0.5 equiv of
solution specific to the analysis of receptor 4 was to use anion C60 in dichloromethane produced a brown solution (Scheme 9).
concentrations sufficient to maintain the cone conformation Jeppesen and co-workers reported a couple of environmentally
when studying the interactions with C60. On the basis of these responsive, supramolecular switching devices107,108 (Schemes 10
controlled continuous-variation studies and solid-state analyses and 11) consisting of receptor 4 and a number of neutral guest
(vide infra), the binding mode, simple 1:1 complexation of the molecules, such as planar or quasiplanar nitroaromatic guests,
C60 within the calix[4]pyrrole cup, is now considered to be including TNB (A), methyl-2,5,7-trinitrofluorenone-4-carbox-
established (Scheme 9). Apart from the complex stoichiometry, ylate (B), and methyl-2,5,7-trinitrodicyanomethylenefluorene-4-
other aspects of the initial study, particularly the visible, carboxylate (C), as well as spherical guest-molecules, including
spectroscopic, and electrochemical findings, are considered fullerene (D), its derivative N-methylfulleropyrrolidine (E), and
correct. These latter findings are discussed below. a snake-like bidentate system consisting of a trinitrodicyanome-
The importance of anion binding to the calix[4]pyrrole thylenefluorene attached to a C60 moiety (F).
framework, a predicate for forming a supramolecular complex To investigate the effects of the individual chemical inputs, C
with the fullerene guest, was recognized at the time of the initial and E were added to the receptor system 4 in both the absence
report. For instance, no significant absorption is seen for receptor and presence of anions. It was observed that, after addition of 2.0
system 4 or C60 guest molecule in dichloromethane solution molar equiv of C (0.8 mM) and 0.5 equiv of E (0.2 mM) to a 0.4
beyond λmax ≥ 550 nm. However, when C60 is added to receptor mM CH2Cl2 solution of receptor 4, the color changed from
4 in the presence of a Cl− anion source (which transforms the yellow to brown almost instantaneously on the laboratory time
1,3-alternate structure to the corresponding bowl-shaped scale (Scheme 10).
conformer), the solution turns green immediately on the This visual color change was thought to be due to the
laboratory time scale as a result of the formation of more formation of the complex C2⊂4, wherein receptor 4 resides in its
ordered multicomponent supramolecular ensemble. The 1,3-alternate conformation. The absorption spectrum of this
absorption spectrum of this complex, which was subsequently presumed complex displayed CT bands centered at λmax = 525,
assigned to C60⊂4·Cl−, is characterized by a strong CT band with 720, and 1220 nm (ε = 3700, 2000, and 2800 M−1 cm−1) (cf.
λmax = 725 nm. In contrast, the addition of 2 equiv of receptor 4 in Figure 12a). Addition of 2 equiv of chloride anions to this
its anion-free 1,3-alternate conformation to a solution of C60 in solution results in a conformational change of receptor 4 from
2654 DOI: 10.1021/acs.chemrev.6b00375
Chem. Rev. 2017, 117, 2641−2710
Chemical Reviews Review

Scheme 8. Initial Suggestion Regarding the Supramolecular Host−Guest Complexation between Receptor 4 and C60

the 1,3-alternate to its cone conformation 4·Cl−. As a spherical-shaped fullerene head and a nearly planar aromatic
consequence, the 2:1 complex C2⊂4 breaks up and is replaced trinitrodicyanomethylenefluorene (TNDCF) tail, and was
by a modified fullerene complex. Initially, the stoichiometry of designed to interact with 4 through two limiting recognition
this complex was assigned as being 1:2 [i.e., E⊂(4·Cl−)2] as modes as shown in Scheme 11. However, subsequent studies of
shown in Scheme 10. Although never reinvestigated in detail, C60 binding106 led to the suggestion that the dominant species
subsequent studies of C60 binding106 led us to propose that the formed upon substrate binding involves a 1:1 receptor/fullerene
species in question is a 1:1 receptor/fullerene complex (Scheme complex (Scheme 13).
12). The conformationally flexible nature of receptor 4 was used to
The anion-induced switching event is readily observed as a follow the complexation of the bifunctional substrate F under
prominent visual color change from brown to green. The conditions that allow for input-based control. In the presence of
absorption spectra for these species are summarized in Figure chloride anion, binding of the C60 head is preferred (because of
12a. Washing with water removes the chloride anions from the the presence of the cone conformer of 4 that favors fullerene
supramolecular fullerene complex and regenerates the initial 1,3- binding). In contrast, complexation of the TNDCF tail
alternate conformer. As a result, the complex C2⊂4 is re- dominates in the absence of a chloride anion source. This is
established. This chemical transformation is reflected in an because, in the absence of anions, receptor 4 exists predom-
observable change in the absorption spectrum (Figure 12a). inantly in the 1,3-alternate conformer, which favors the binding
Addition of a new aliquot of chloride anions to the CH2Cl2 phase of planar electron-deficient aromatic guests.
serves to regenerate the green color, as would be expected given These two binding modes were reflected in (i) the formation
the presumed regeneration of the fullerene complex (cf. Figure of two different CT complexes and (ii) a change in the optical
12a). These experiments provide additional support for the signature of the system that could be monitored qualitatively by
notion that the differential recognition properties of receptor 4 the naked eye, by steady-state absorption spectroscopy (Figure
can be controlled by adding or removing chloride anions 12c), and more quantitatively by time-resolved flash photolysis
(“input”) and that the specific chemical “lead” (i.e., the nature of studies. Therefore, this system functions as an environmentally
the guest) has a demonstrable effect on the optical “output” from responsive supramolecular switching device, wherein the ON
the overall “device” (i.e., the color of the solution). and OFF states are controlled by the chloride anion
An interesting binding mode was also demonstrated with the concentration and the effect of switching is “read out” in terms
snake-like trinitrodicyanomethylenefluorene-attached-C60 (F) of the observable changes in the color and spectroscopic
derivative. This species contains two different binding motifs, a properties.
2655 DOI: 10.1021/acs.chemrev.6b00375
Chem. Rev. 2017, 117, 2641−2710
Chemical Reviews Review

Scheme 9. Revised View of the Supramolecular Interactions between Receptor 4 and C60 Involving Stabilization of a 1:1 Complex
as the Dominant Species

Tetraalkylammonium cations can compete with small full- calix[4]pyrrole anion complexes (cf. Figure 13a). In retrospect,
erenes as substrates for TTF-calix[4]pyrroles. Analyses of this it is likely that this competition contributed in part to the original
effect led to an appreciation that small fullerenes bind to TTF- receptor/fullerene binding stoichiometry being assigned in-
calix[4]pyrroles with a 1:1 stoichiometry under most conditions, correctly as 2:1.
as opposed to the 2:1 binding stoichiometry initially 2.1.4. TTF-Annulated Calix[4]pyrroles and Surface
proposed.105,107,108 The most detailed study was carried out Studies. Sanguinet and co-workers reported on the self-
using receptors 4 and 5.106 This study included C70 as well as C60 assembled monolayer (SAM) systems constructed by mono-
and a number of chloride anion salts. On the basis of both TTF-calix[4]pyrroles, functionalized with an alkane-thiol
solution-phase and solid-state structural analyses, it was anchoring group (10 and 11), for anion sensing on solid-
concluded that both C60 and C70 are bound within the bowl- supported devices.109 The synthetic strategy used to obtain 10
like cup of the anion-complexed cone conformers of these TTF- and 11 is outlined in Scheme 15. The two SAMs fabricated using
calix[4]pyrroles in a ball-and-socket binding mode (Scheme 14 10 and 11 were designed to permit anion complexation to be
and Figure 13) in the absence of a competing cationic guest. The detected by binding-induced changes in the electrochemical
binding affinities of receptor 5 for C60 and C70 could be properties of the TTF units. Two monopyrrolo-TTF references
modulated through the choice of the coordinating anion, namely, (MPTTF1 and MPTTF2) were also synthesized to act as
F−, Cl−, and Br−. All three of these anions bind to 4 and 5 and controls. A comparison of the surface coverages obtained for the
stabilize the cone conformer of the TTF-calix[4]pyrrole. Thus, SAMs prepared from the two model compounds MPTTF1 and
all three act as positive heterotropic allosteric regulators for the MPTTF2 with those obtained for the SAMs derived from the
complexation of C60 and C70 by 4 and 5. TTF calix[4]pyrroles 10 and 11 revealed that the MPTTF model
However, the specific choice of anion has a strong effect on the compounds provided coverages that were almost 3 times higher
thermodynamic stabilities of the resulting supramolecular [i.e., (1.9 ± 0.2) × 10−10 and (0.7 ± 0.1) × 10−10 mol cm−2 for
assemblies, as inferred from the fullerene binding constants MPTTF1 and 10, respectively]. Based on Cory−Pauling−
recorded in dichloromethane solution. In fact, the selectivity of Koltun (CPK) models, the areas per molecule for MPTTF1 and
receptor 5, that is, whether C60 or C70 was bound preferentially, 10 were estimated to be 49 ± 7 and 79 ± 11 Å2, respectively. On
was found to be a function of the halide anion used to stabilize the this basis, it was concluded that the SAMs constructed using the
cone conformer of the TTF-calix[4]pyrroles. Moreover, the TTF-calix[4]-pyrroles 10 and 11 were not organized or as well
nature of the tetraalkylammonium countercation plays a crucial packed on the monolayer surface as those prepared using the
role in the binding process of C60 and C70. In particular, smaller control MPTTF model compounds.
cations, such as tetraethylammonium (TEA+) and, to a lesser CV analyses were used to probe the changes in the redox
extent, TBA+, compete with the fullerenes for encapsulation potentials of the SAMs prepared from compounds 10 and 11,
within the host cavity (Figure 13a). As a result, they act as along with the reference systems, as a function of Cl −
antagonists for one another. Solid-state findings consistent with concentration under solution-phase conditions (dichlorome-
this competition included both an early crystal structure (cf. thane). Although it was expected that addition of Cl− to the
Figure 11a) and a number of later X-ray structures of SAMs derived from the two MPTTF model compounds would
2656 DOI: 10.1021/acs.chemrev.6b00375
Chem. Rev. 2017, 117, 2641−2710
Chemical Reviews Review

the addition of Cl− to a solution of the TTF-C[4]P-based system


11 (Δ E1/21 = −170 mV) and to the SAM derived from it (ΔE1/21
= −30 mV) revealed that the Cl−-induced shift was much smaller
in the SAM system than in solution. This observation is ascribed
to the fact that immobilization of the TTF-calix[4]pyrrole
receptor renders more difficult the conformational change from a
1,3-alternate conformation to the cone conformation that is a
known predicate to anion binding. In fact, it is possible that only a
fraction of the immobilized TTF-C[4]Ps have enough free space
to undergo this crucial conformational change and thus act as
anion-binding and signaling sites. Support for this latter notion
came from a close inspection of the cyclic voltammograms
recorded in the presence of Cl−. The cyclic voltammograms
show (Figure 14) that the oxidation and reduction peaks became
broader upon addition of Cl−. Such an observation is consistent
with a system that results from the superposition of two different
redox processes, one corresponding to the first redox process
associated with the uncomplexed TTF-C[4]P 11 and one
corresponding to the complex (i.e., 11·Cl−).
2.1.5. TTF-Annulated Calix[4]pyrroles for Chemores-
ponsive Supramolecular Copolymerization with Hetero-
complementary Calix[4]pyrroles. A new set of artificial self-
assembled materials,110 constructed by mixing two different
Figure 11. (a) Single-crystal X-ray structure of 4 and C60 wherein heterocomplementary macrocyclic receptors, namely, the
outside guest binding is seen in the solid state. (This figure was redrawn electron-rich TTF-calix[4]pyrroles (4−6) and electron-deficient
using data that were originally published in ref 105.) As detailed in the bis(dinitrophenyl) meso-substituted calix[4]pyrroles (12 and
text, a 1:1 encapsulated complex is now believed to pertain in solution. 13) (Figure 15), was reported by Sessler et al.
Note that the disordered propyl substituents on the TTF subunits have The resulting supramolecular polymeric materials, stabilized
been truncated for clarity. (b) Vis−NIR spectra of (i) a 0.135 mM
solution of 4, (ii) a 0.0687 mM solution of C60, (iii) the putative complex
by a combination of donor−acceptor-induced π−π-stacking and
with fullerene produced after the addition of 0.5 equiv of C60 to the hydrogen-bonding interactions, undergo dynamic, reversible
chloride-anion bound form of 4, (iv) a mixture of 4 with 0.5 equiv of C60, dual-guest-dependent structural transformations upon exposure
and (v) a mixture of 4 and 0.5 equiv of C60 in the presence of excess to different types of external chemical inputs, namely, Cl− and
TBACl. All spectroscopic measurements were carried out in CH2Cl2 at TNB. The supramolecular polymerization process and dynamics
298 K. (Reproduced with permission from ref 105. Copyright 2006 of the copolymers along with their analyte-dependent responsive
Wiley-VCH Verlag GmbH.) behavior were established by various spectroscopic techniques.
Most significantly, it was demonstrated that the copolymer
Table 5. Reduction Potentials of C60 Observed Before and formed from the supramolecular interactions between the
After the Addition of a Cl− Anion Salt, Receptor 4, and ditopic monomers collapsed to its nonoligomerized species
Receptor 4 in Combination with a Cl− Anion Source at 298 K upon the addition of either Cl− or TNB. Under solution-phase
in CH2Cl2 conditions, dynamic light scattering (DLS) and diffusion-
entity E11/2 (V) E21/2 (V) ΔE11/2 (V) ΔE21/2 (V)
ordered NMR (DOSY-NMR) spectroscopy were used to
establish the formation of deaggregated species and the
C60 −1.444 −1.045 − −
formation of solutions with lower net viscosity, respectively.
C60 + Cl− −1.440 −1.042 +0.004 +0.003
Solid-state structural information was obtained for both the
C60 + 4 −1.443 −1.056 +0.001 −0.011
monomeric and polymeric materials.
C60⊂(4·Cl−) −1.437 −1.119 +0.007 −0.074
a
Sessler and co-workers reported another self-assembled
All measurements were performed in anhydrous deoxygenated system111 constructed from 5 and an appropriately chosen
CH2Cl2 solution, at ambient temperature and under an Ar atmosphere. calix[4]pyrrole pyridinium salt, either 15 or 16 (Figure 16).
The concentration of C60 was 0.2 mM and 0.10 M of TBAPF6 was
added as supporting electrolyte. Cyclic voltammograms were obtained
When the noncoordinating BArF − {tetrakis[bis(3,5-
with a sweep rate of 200 mV s−1. The half-wave potentials E1/2 trifluoromethyl)phenyl]borate} anion salt 15 was used, a linear
reported were obtained from an average of the cathodic and anodic supramolecular oligomer was formed upon mixing with 5 in a 1:1
cyclic voltammetric peaks. All potentials are referenced to the ratio in chloroform. A very different species, a structurally
ferrocene/ferrocenium couple (Fc/Fc+) characterized capsule product, was obtained when the
corresponding iodide salt 16 was used (cf. Figure 16 and
lead to only minor negative shifts in the first oxidation waves, it discussion below).
was actually found that the addition of a large amount of Cl− to The supramolecular ensembles produced from mixtures of 5
the SAMs induced a small but significant positive shift (ΔE1/21 = and 15 or 5 and 16 were found to be dependent on the nature of
+15 mV). This observation, which remains to be explained, the salts present in solution, as inferred from studies with
stands in marked contrast to what was seen in the case of the tetraethylammonium tetrakis[bis(3,5-trifluoromethyl)phenyl]-
SAMs derived from 10 and 11. In these cases, the progressive borate (TEABArF), tetrabutylammonium iodide (TBAI), and
addition of Cl− (from 0 to 0.5 μm) induced a significant negative tetraethylammonium iodide (TEAI). In fact, mixtures of 5 and
shift (ΔE1/21 ≈ −30 mV) in the first oxidation process (Figure 15 were found to give rise to three different supramolecular
14). A comparison of the maximum negative shifts induced by complexes from one parent system, (5·15)n, upon addition of
2657 DOI: 10.1021/acs.chemrev.6b00375
Chem. Rev. 2017, 117, 2641−2710
Chemical Reviews Review

Scheme 10. Binding Modes Initially Proposed for Receptor 4 Induced by Different Inputs (A−E) in the Presence and Absence of a
Chloride Anion Source

a
On the basis of subsequent experiments, the binding stoichiometry for the fullerene complex is now considered to be 1:1. See Scheme 12 and ref
106.

Scheme 11. Binding Modes Initially Proposed for the Interaction of Receptor 4 and the Bifunctional Substrate F in the Presence
and Absence of Chloride Anion

2658 DOI: 10.1021/acs.chemrev.6b00375


Chem. Rev. 2017, 117, 2641−2710
Chemical Reviews Review

conformational switch to give its iodide-bound cone conformer.


Under the experimental conditions employed, the tetrabuty-
lammonium cation was not encapsulated within the bowl-like
cavity of the resulting anion complex, [5·I]−. Instead, it was
calix[4]pyrrole 15 that was bound in a 2:1 capsule-like fashion to
give 15⊂2[5·I]−, as shown in Figure 16.
Sessler and co-workers also reported another short supra-
molecular dynamic oligomeric system.112 This environmentally
responsive material was produced from two heterocomplemen-
tary subunits, namely, glycol diester-linked bis-2,5,7-trinitrodi-
cyanomethylenefluorene-4-carboxylate (TNDCF) and 4. The
underlying interactions are shown schematically in Figure 17.
When both of these components are mixed in organic solvents
(e.g., CHCl3, CH2ClCH2Cl, or methylcyclohexane), supra-
molecular aggregation takes place to produce supramolecular
oligomers, (4·G)n, that are presumably stabilized by hydrogen-
bonding and donor−acceptor-based CT interactions. The
Figure 12. (a,b) UV−vis−NIR spectral changes observed for a fluorescence emission of the TNDCF subunit is quenched
“switching device” based on the TTF-calix[4]pyrrole 4 observed in under conditions that promote the formation of supramolecular
the absence and presence of various inputs in CH2Cl2 at room
aggregates containing 4 and TNDCF. However, it is restored
temperature. (c) Visual effects under ambient conditions. (This figure
was updated based on the findings presented in Schemes 12 and 13. The when the oligomer is destroyed, leading to the suggestion that
original version with errors in complex stoichiometry originally oligomer (4·G)n could act as a dual-analyte chemoresponsive
appeared in ref 107. Copyright 2008 American Chemical Society.) sensing system. In fact, the combination of quenching and
analyte-induced “turn-on” fluorescence was used to test for
each of these inputs. The noncoordinating nature of the BArF− anions and nitroaromatic explosives in initial proof-of-concept
anion means that the addition of the TEA+ cation in the form of experiments (Figure 17).
its BArF− salt (TEABArF) to a mixture of 5 and 15 in chloroform To understand further the determinants that control self-
does not trigger any conformational change in either of the two association in calix[4]pyrrole-based systems, Sessler and co-
calix[4]pyrrole units. Therefore, there is no calix[4]pyrrole cavity workers recently reported the results of a study wherein the rigid
into which the TEA+ cation can be encapsulated. calix[4]pyrrole systems 12, 13, 15, and 16 used initially were
On the other hand, the pyridinium subunits present in the Pyr- replaced by a set of new flexible alkyl-linked bis-dinitrophenyl
C[4]P dication 15 can “intercalate” into the “clefts” that esters (i.e., H−J, Figure 18).113
characterize the 1,3-alternate form of 5. This results in self- Mixing 4 and any of the esters H−J in chloroform gave rise to
assembly and the production of a linear aggregate, (5·15)n, an immediate color change on the laboratory time scale, with the
wherein the noncoordinating BArF− anions provide for charge solution turning from yellow to green. This color change was
balance. When TEABArF was replaced by TBAI in the above accompanied by an increase in the absorbance intensity at 566
studies, the electron-rich BTTF-calix[4]pyrrole 5 underwent a nm in the UV−vis−NIR absorption spectrum, a finding

Scheme 12. Dominant Binding Modes Seen for Receptor 4 Observed with Different Substrates (A−E) and in the Presence and
Absence of a Chloride Anion Source

2659 DOI: 10.1021/acs.chemrev.6b00375


Chem. Rev. 2017, 117, 2641−2710
Chemical Reviews Review

Scheme 13. Dominant Binding Modes Observed for Receptor 4 and the Bifunctional Substrate F in the Presence and Absence of
Chloride Anion

Scheme 14. Proposed Binding Mode for Receptor 4 and C60


or C70 as Observed in the Presence of Halides and (Inset)
Single-Crystal Structure Showing 1:1 Binding Mode with the
Same Net (TBA)[C60⊂5·F] Stoichiometry as Believed To Be
Operative in Solution

Figure 13. Single-crystal X-ray diffraction structures of (a) [TEA⊂(5·


Cl)], (b) (TOA)[C60⊂(4·Cl)], (c) (TOA)[C70⊂(4·Cl)], and (d)
(THA)[C70⊂(5·Cl)]. For clarity, the C60 and C70 moieties are depicted
as violet and purple space-filling models, respectively. Note: TEA+ =
tetraethylammonium cation, TOA+ = tetraoctylammonium cation, and
THA+ = tetrahexylammonium cation. (This figure was redrawn using
data that were originally published in refs 106 and 110. Copyright 2014
American Chemical Society and 2011 National Academy of Sciences,
respectively.)

the proposal that all three combinations of 4 with H−J gave rise
to self-associated species in chloroform. Dilution-based
absorption spectroscopic measurements carried out in chloro-
considered diagnostic of a CT interaction and, hence, self- form of the self-assembled systems revealed Ka values of 1.5 ×
association between the species. Evidence for stabilizing 103, 1.3 × 103, and 3.8 × 103 M−1, respectively, corresponding to
hydrogen-bonding interactions and through-space proton- the 1:1 binding constants leading to (4·H)n, (4·I)n, and (4·J)n.
coupling interactions between 4 and the dinitrophenyl moiety Another supramolecular oligomer constructed by two redox-
of the bis-esters came from 1H NMR spectroscopic analyses. X- responsive heteroditopic monomers, 4 and K, was reported in
ray structures, revealing the presence of linear, self-assembled 2015.114 This more complex system utilizes two different binding
polymeric entities in the solid state, provided further support for modes, where the carboxylate of [6,6]-phenyl-C61-butyric acid
2660 DOI: 10.1021/acs.chemrev.6b00375
Chem. Rev. 2017, 117, 2641−2710
Chemical Reviews Review

be disaggregated through the addition of an organic acid (e.g.,


methanesulfonic acid, MSA) or electrolysis (Figure 19). A
noteworthy feature of this system is that it can be controlled
through the use of highly orthogonal inputs, namely, chemical
and electrochemical, which leads to the suggestion that more
complex systems, such as this one, might have a role to play in the
creation of stimulus-responsive systems where external factors
are used to modulate the structure and function of self-assembled
materials over a range of length scales.
2.1.6. TTF-Annulated Calix[4]pyrroles for Ion-Medi-
ated Reversible Electron-Transfer (ET) Process in Supra-
molecular Ensembles Formed with Electron-Deficient
Guests. Anion-mediated reversible ET in a synthetic host−guest
assembly that mimics aspects of cofactor-controlled biological
electron transfer was reported in 2010 by Sessler and co-
workers.115 As with other calix[4]pyrrole ensembles, anion
Figure 14. Cyclic voltammograms of self-assembled monolayers
prepared from compound 11 measured in dichloromethane (surface
binding to 4 in this instance produces a cone conformer that can
coverage of 11 was estimated to be 0.7 ± 0.1 × 10−10 mol cm−2) act as a receptor for large cationic guests, namely, the electron-
recorded after additions of successive aliquots of TBACl. A Fc/Fc+ deficient benzimidazolium quinone (BIQ2+). The resulting
reference electrode was employed using 0.1 M TBAPF6 as the complex undergoes intraensemble thermal electron transfer to
supporting electrolyte. A 0.2 V s−1 scan rate was used. Inset: CV produce a long-lived charge-separated (CS) state. Based on X-ray
titration curves showing the negative shift of the first oxidation potential structural analysis, this supramolecular capsule is best described
(measured at the maximum intensity) of 11 seen upon addition of Cl− as a tightly coupled biradical species, [4]2•+·BIQ•+·2Cl−, wherein
ions. (Reproduced with permission from ref 109. Copyright 2009 Wiley- two cone-like anion-bound conformers of 4 (one of which is
VCH Verlag GmbH.)
oxidized by one electron) serve to encapsulate one BIQ•+ guest
(Figure 20).
(PCBA) binds through hydrogen bonding to the pyrrole NH As noted above, anion binding to 4 triggers conversion to the
protons of 4 in its cone conformer and the spherical C61 subunit cone conformer and favors forward ET to the BIQ2+ acceptor. In
present in PCBA is bound in the cavity of 4 through CT contrast, the addition of a tetraethylammonium cation source,
interactions. The resulting ensemble could be modulated by which serves to compete for the “pocket” of 4·Cl−, leads to
means of chemical and electrochemical stimuli. Addition of an electron back-transfer from BIQ•+ to 4•+. This restores the initial
organic base (e.g., 1,8-diazabicyclo[5.4.0]undec-7-ene, DBU) oxidation states of the donor−acceptor pair (Figure 21). This
initiates self-assembly of the monomers through a molecular type of ON−OFF switchable ET behavior can be observed by
switching event. The resulting supramolecular ensemble could monitoring the changes in absorption intensity at 751 and 1995

Scheme 15. Synthetic Route to TTF-Calix[4]pyrroles with Alkanethiol Anchoring Groups Used to Prepare SAMs

2661 DOI: 10.1021/acs.chemrev.6b00375


Chem. Rev. 2017, 117, 2641−2710
Chemical Reviews Review

Figure 15. Schematic representation of the supramolecular self-assembly between the TTF-calix[4]pyrroles 4, 5, and 6 and the heterocomplementary
dinitrophenyl-calix[4]pyrroles (DNP-C[4]P) 12 and 13. Chemoresponsive behavior is seen upon the addition of either TEACl or TNB. (Reproduced
with permission from ref 110. Copyright 2011 National Academy of Sciences.)

from receptor 4 than the native fullerene. After treating host 4


with Li+@C60 and a Cl− anion source in the absence of a
competing cation, ET is switched ON to produce a 1:1 radical ion
pair between 4•+ and [Li+@C60]•− with a net stoichiometry of
Li+@C60⊂4·Cl−. Addition of tetraethylammonium chloride
provides a source of TEA+, a species that acts as a competitive
inhibitor for the neutral Li+@C60•− ET product and its cationic
precursor, Li+@C60. As a consequence, these fullerene species are
released from the calix[4]pyrrole cavity, resulting in electron
back-transfer and restoration of the initial donor and acceptor
pair. Evidence in support of these conclusions is summarized
briefly below.
The appearance of an absorption band at 1035 nm upon the
gradual addition of tetra-n-hexylammonium chloride (THACl)
to a benzonitrile solution of 4 and Li+@C60 is ascribed to the
generation of the one-electron-reduced species [Li+@C60]•−, as
shown in Figure 23a. The intensity of this peak gradually
increases upon the addition of THACl until it reaches a constant
value, at which point the ET process is considered essentially
complete. Upon addition of TEACl to this supramolecular ion-
pair system, Li+@C60•−⊂4•+·Cl−, the absorbance feature at 1035
nm gradually decreases in intensity, a finding consistent with the
breakup of the radical-ion-pair complex and generation of the
new supramolecular system TEA+⊂4·Cl−.
Further evidence for the formation of a radical-ion-pair
complex came from EPR analyses of the species produced after
Figure 16. Schematic representation of three limiting equilibrium states the proposed ET from 4 to Li+@C60. When the spectrum was
that can be produced from 5 and 15 depending on the nature of the recorded in benzonitrile at 298 K (Figure 23b), a four-line signal
experimental conditions employed. In the absence of a coordinating was observed. This signal was ascribed to a 4•+·Cl− radical cation,
anion or competing cation source, the self-assembled oligomer (5·15)n in which the electron spin is localized on only one TTF unit,
is obtained in CHCl3. Changes in this oligomer can be triggered by the
addition of various tetraalkylammonium salts. (Reproduced with
rather than delocalized throughout a π-dimer radical-cation
permission from ref 111. Copyright 2013 American Chemical Society.) complex involving another TTF unit. The signal at g = 4.43
(Figure 23c) was attributed to the interaction between the two
unpaired electrons, which generates a triplet state. A single-
nm in the steady-state absorption spectra, as well as by electron crystal structure of the proposed radical ion pair was obtained; it
paramagnetic resonance (EPR) spectroscopy (Figure 22). is shown in Figure 24a,b.
Another supramolecular complex involving host system 4 that Subsequently, an example of photoinduced electron transfer
undergoes ground-state ET was reported in 2011.116 This system (PET) from an electron-rich TTF-calix[4]pyrrole, 5, to a
was constructed using Li+@C60, a species that, like BIQ2+, is porphyrin carboxylate, PA, within a supramolecular ensemble
complexed by the anion-bound cone conformer of 4. Compared was reported.118 A carboxylate moiety was chosen as the anion
to pristine C60, Li+@C60 is a more powerful electron acceptor.117 moiety in constructing the proposed supramolecular complex
This makes it a more promising candidate for ground-state ET involving 5 and PA because carboxylate anions are known to bind
2662 DOI: 10.1021/acs.chemrev.6b00375
Chem. Rev. 2017, 117, 2641−2710
Chemical Reviews Review

Figure 17. Graphical representation of the proposed supramolecular ensemble formed from 4 and TNDCF (G) and schematic view of the breakup of
the supramolecular oligomer that occurs upon exposure to TNB (left) or a chloride anion source (right). In both cases, fluorophore G is released.
(Reproduced with permission from ref 112. Copyright 2016 Wiley-VCH Verlag GmbH.)

to the NH protons of the calix[4]pyrroles in a 1:1 fashion and to form. The sum total of the findings provides support that the
induce a conformational change from the 1,3-alternate to the three components assemble to produce a multicomponent
cone conformer in analogy to what is seen in the case of the redox-active ensemble.
chloride anion. In fact, the PA anion (TEA+ countercation) was 2.2. TTF-Annulated Hybrid Calix[n]pyrrole Systems
found to bind to 5 with a 1:1 stoichiometry (Figure 25a) and a
binding constant of 6.3 × 104 M−1 at 298 K in benzonitrile. 2.2.1. “Pacman”-Type Schiff-Base TTF-Calix[4]pyrrole.
In initial studies, TEA+ was chosen as the countercation to the A very different class of flexible TTF-pyrrole-derived macrocyclic
carboxylate moiety, as it is known to bind strongly inside the molecular receptor is the Schiff-base system 17, reported by
bowl-shaped cavity of the cone conformer of TTF-calix[4]- Sessler and co-workers.120 This so-called Pacman receptor
system combines the complexation and physiochemical proper-
pyrroles.106 This dual binding to receptor 5, involving both TEA+
ties of the benzoannulated-TTF pyrrole (BTTF-pyrrole) used to
and the carboxylate moiety, was expected to enhance the
construct 5 with the synthetic versatility and simplicity of Schiff-
interaction between the porphyrin PA and 5, leading to stronger
base condensation procedures. The flexible framework of the
supramolecular complexation. Laser photoexcitation of this
Pacman receptor and its Schiff-base construction is thought to
supramolecular complex resulted in the formation of the triplet underlie its ability to stabilize a number of metal complexes. The
charge-separated state composed of a TTF•+ radical cation and synthetic route leading to 17 is summarized in Scheme 16. Its
the radical anion PA•−. The rate constants corresponding to the metal complexation chemistry is discussed below.
forward and backward intramolecular electron transfer within the Receptor 17 was found to react with a Pd(II) source to
ensemble were determined to be 2.1 × 104 and 3.6 × 102 s−1, produce a bimetallic complex 17·Pd2. The formation of this
respectively. The rate constants of intermolecular forward and complex was found to help in the stabilization upon oxidation of
backward electron transfer occurring in the absence of complex an otherwise difficult-to-access mixed-valence TTF radical
formation were also determined and found to be 4.4 × 108 and dimer. EPR and optical spectroscopies were used to characterize
9.8 × 108 M−1 s−1, respectively. the mixed-valence species. The crystal structure of complex 17·
A rare example of a three-component supramolecular Pd2 supports its Pacman-like structure (Figure 27).
ensemble undergoing PET was also reported (Figure 25b).119 Sessler and co-workers also reported a TTF-based macrocyclic
In this case, Li+@C60 was used in conjunction with 5 and a host system, the TTF-calix[2]pyrrole[2]thiophene 18,121 which
different porphyrin carboxylate anion, PA2. Photoexcitation of can bind electron-deficient planar guest molecules in a Pacman-
the three-component mixture resulted in PET from the triplet like fashion. The system acts as a chemosensor in that a readily
excited state of PA2 to 5•+. This produced a higher-energy visible color change is produced in the presence of nitro-
charge-separated state that was proposed to consist of the aromatics, such as TNB and TNP. The color presumably reflects
supramolecular radical ion complex Li+@C60•−⊂5·PA2•+, which formation of a D−A-type complex that is stabilized as the result
decays to the ground state with a lifetime of 4.8 μs. A of π−π donor−acceptor interactions. The synthesis of 18 is
characteristic signal at 1035 nm in the vis−NIR absorption outlined in Scheme 17.
spectra (Figure 26a) is consistent with the formation of a charge- Solid-state structural analysis of this host−guest supra-
separated species, 5•+/Li+@C60•−, where the PA2 is in its neutral molecular system by single-crystal X-ray diffraction established
2663 DOI: 10.1021/acs.chemrev.6b00375
Chem. Rev. 2017, 117, 2641−2710
Chemical Reviews Review

centered at g = 2.010, which is in the region expected for both a


TTF radical cation and a TCNQ radical anion.
2.3. TTF-Annulated Expanded Calix[n]pyrroles
In 2012, Sessler and co-workers reported the synthesis of several
expanded calix[n]pyrroles (n = 5 and 6) containing a higher
number of TTF-pyrroles (compounds 20−22; Scheme 19).123
These systems were of intrinsic interest since they have the same
empirical formula as their known tetrakis-TTF-calix[4]pyrrole
analogues (i.e., 4 and 5), but were expected to display different
conformational features and to have larger cavity sizes. These
new congeners of 4 and 5 were synthesized using the same acid-
catalyzed condensation reaction conditions95,97,98 originally
developed to obtain the TTF-calix[4]pyrrole products. Interest-
ingly, when BTTF-pyrrole was used as a precursor, the expanded
calix[n]pyrroles 20 and 21 were produced, along with 5, in good
to moderate yields. In contrast, when the analogous reaction was
carried out with PrS-TTF-pyrrole, the isolated products
consisted of 4 and 22, with only a trace amount of the putative
hexakis derivative being observed in the matrix-assisted laser
desorption ionization time-of-flight (MALDI-TOF) mass
spectrum of the crude reaction mixture.
Single crystals of the benzo-fused calix[n]pyrrole products 5,
20, and 21 suitable for X-ray diffraction analysis were obtained in
the form of the corresponding chloride anion complexes (as the
TBA+ salts). The resulting structures allowed direct visualization
of the three-dimensional arrangement of the TTF units on the
periphery of each macrocyclic core in the solid state (Figure 30).
It also allowed some inferences about the anion-binding features
to be drawn.
Figure 18. Single-crystal X-ray structure of the proposed supra- Within the series consisting of 5, 20, and 21, 1:1 chloride anion
molecular assembly formed when 4 and the bis-dinitrophenyl esters H− complexes are seen in the solid state. In the case of receptor 21,
J are mixed in noncompetitive organic media. (Reproduced with all six pyrrolic protons interact with the bound chloride, as
permission from ref 113. Copyright 2014 Royal Society of Chemistry.) inferred from the structural parameters (Cl···H−N distances of
2.39 ± 0.01 Å). The Cl− anion is centered in the cavity and
resides within an imaginary plane defined by the hexagonal
arrangement of the pyrrolic nitrogen atoms. In contrast, four
a “cleft-like” arrangement wherein the two TTF units lie parallel NH···Cl− hydrogen-bonding interactions were seen in the case of
to one another and are separated by approximately 3.44−3.56 Å. receptors 5 and 20, as judged from the Cl···HN distances (2.49 ±
The structure in question is shown in Figure 28. This distance 0.07 Å for 5 and 2.43 ± 0.01 Å for 20). On this basis, it was
allows for the formation of Pacman-type sandwich complexes concluded that receptors 5 and 20 are not as optimized for Cl−
with planar electron-deficient guest molecules.
binding as 21.
2.2.2. TTF-Annulated Calix[2]pyrrole[2]thiophene. In
Analysis of the binding isotherms obtained from UV−vis
work that bears some analogy to that discussed above, Jeppesen
spectroscopic titrations carried out in CHCl3 in the presence of
and co-workers reported a rare example of a TTF-calix[2]-
pyrrole[2]thiophene hybrid system, 19, wherein two thiophene both anionic (Cl−, Br−, I−, MeCO2−, H2PO4−, and HSO4−, as the
moieties are incorporated within the central macrocyclic skeleton TBA+ salts) and neutral (TNB and TNT) substrates revealed
(Scheme 18).122 that the hexpyrrolic system 21 is the most efficient molecular
Keeping with the analogy between 18 and 19, the latter system receptor for anions (e.g., Ka = 5.3 × 106 M−1 in the case of Cl− as
was found to act as a molecular receptor for TCNQ, which was its TBA+ salt; see Table 6). In contrast, compound 5 proved most
bound in a Pacman-like arrangement. The guest-binding ability effective for the recognition of electron-deficient nitroaromatic
of 19 was supported by single-crystal X-ray diffraction studies explosives, such as TNB and TNT (e.g., Ka = 1.7 × 104 M−1 in the
(Figure 29). On the basis of this structure, guest binding is case of TNB). The relatively high anion affinities of 21 were
stabilized by CT interactions and overlap of the π-surfaces of the ascribed to the size matching revealed in the structural analysis
two TTF side arms with the electron-deficient TCNQ. discussed above, as well as the presence of six presumably
Support for the presence of CT interactions came from UV− preorganized NH···X− hydrogen-bonding interactions. The fact
vis−NIR spectroscopic studies, as well as EPR studies using that 5 was the best receptor in the series for both TNB and TNT
CH2Cl2 as the solvent. Neither 19 nor TCNQ gives rise to was rationalized in terms of the TTF substituents being able to
significant absorption features beyond 700 nm. However, upon adopt conformations suitable for “sandwich-type” substrate
mixing of these components, two newly generated CT bands recognition through D−A interactions, as well as the attendant
appear at λmax = 749 and 851 nm. These optical features are cooperative effects seen for planar electron-deficient substrates in
characteristic of a TTF-to-TCNQ CT interaction. The EPR the case of the TTF-calix[4]pyrroles discussed above. Both of
spectrum also revealed the presence of a weak radical signal these effects rely on an ability to access well-defined 1,3-alternate
2664 DOI: 10.1021/acs.chemrev.6b00375
Chem. Rev. 2017, 117, 2641−2710
Chemical Reviews Review

Figure 19. (a,b) Chemical structures of the heteroditopic monomers (a) 4 and (b) K. (c) Illustrations of an anion-bound form of the 4 and the phenyl
C61 butyric acid (PCBA, K). (d) Schematic representation of the self-assembled structure produced from the cone conformer of 4 and the deprotonated
form of K. Also shown is an approach for controlling this system through acid−base chemistry and electrochemical oxidation. (Reproduced with
permission from ref 114. Copyright 2015 American Chemical Society.)

properties to be further tuned. Axial ligand coordination to the


metal center might provide an additional specific approach to
modulating the features of these systems. These possibilities
make the marriage of TTF and porphyrin chemistry attractive; it
might offer a range of potential supramolecular D−A systems
useful for a variety of prospective applications.
One way to incorporate TTF into conjugated pyrrolic
frameworks would be through direct or stepwise condensation
with suitable aryl aldehydes. In fact, using these and related
approaches, a number of TTF-annulated porphyrins, specifically
23−31, containing differing numbers of fused TTF units have
been prepared.124−127 Their chemical structures are shown in
Chart 3. These systems are discussed further below.
In some cases, a simple and straightforward acid-catalyzed
macrocyclization protocol involving the condensation of a TTF-
Figure 20. X-ray crystal structure of the Cl−-induced supramolecular annulated pyrrole with an electron-deficient aldehyde under
radical-ion-pair capsule constructed by 4 and BIQ2+ of net stoichiometry dilute conditions results in the effective formation of TTF-
[4]2•+·[BIQ] •+·2Cl−. (Reproduced with permission from ref 115. functionalized porphyrin derivatives. In fact, this approach was
Copyright 2010 American Association for the Advancement of Science.) used to prepare compounds 28−31. However, to synthesize
mono- and bis-TTF-annulated porphyrin derivatives (23−26),
molecular conformations, something that is possible only for 5 functionalization of the TTF-annulated pyrroles or other
within this congeneric series. precursors proved necessary prior to carrying out the final
2.4. TTF-Annulated Porphyrins macrocyclization reaction. Typically, this functionalization has
2.4.1. TTF-Porphyrins with Different Numbers of Fused taken the form of creating either activated pyrroles or di- or
TTF Units. As mentioned earlier, TTF-annulated pyrroles tripyrrolic precursors. For instance, condensation of TTF-
function as good electron-donor units when incorporated into pyrrole bis-carbinol (TTFPBIOL) with a tripyrrane (TP2 or
anion-binding motifs such as calix[n]pyrroles. In a view of their TTFTP) was used to synthesize 23 and 26, whereas
intriguing redox properties, TTF-pyrroles are also attractive condensation of a TTF-pyrrole with pyrrole bis-carbinol
building blocks for preparing π-conjugated oligopyrrolic macro- (PBIOL) was used to prepare 25. The tris-TTF-porphyrin
cyclic systems (e.g., porphyrins and expanded porphyrins) that (30) was obtained through the condensation of a TTF-pyrrole
might act as photo- or redox-active chromophores. Unlike TTF- with an aldehyde precursor while controlling the molar ratios;
annulated calix[n]pyrroles, these TTF-porphyrin analogues subsequent careful purification by column chromatography
would be expected to form stable metal complexes, which yielded the desired product. The synthetic routes used to obtain
might allow their intrinsic photophysical and electrochemical representative TTF-porphyrins are summarized in Scheme 20.
2665 DOI: 10.1021/acs.chemrev.6b00375
Chem. Rev. 2017, 117, 2641−2710
Chemical Reviews Review

Figure 21. Chemical structures of the molecules, ions, and self-assembled structures associated with the proposed supramolecular ion-mediated ET
process involving receptor 4 and BIQ2+. (Reproduced with permission from ref 115. Copyright 2010 American Association for the Advancement of
Science.)

Figure 22. (a) Results of a UV−vis−NIR spectroscopic titration wherein 4 (30 mM) and 1 molar equiv of BIQ2+2PF6− were treated with increasing
quantities (up to 10 equiv) of THACl in CHCl3. (Inset) EPR spectra of a 1:1 mixture of 4 and BIQ2+2PF6− (1.1 × 10−4 M in each) recorded in CHCl3 at
room temperature upon the gradual addition of increasing quantities of THACl. (b) Cation-induced reverse ET seen upon the addition of up to 15 equiv
of TEACl to a 1:1 solution of 4 and BIQ2+2Cl− (30 mM in each). (Inset) EPR spectra of a 1:1 mixture of 4 and BIQ2+2Cl− (1.1 × 10−4 M in each)
recorded in CHCl3 at room temperature upon the addition of up to 5 molar equiv of TEACl. (Reproduced with permission from ref 115. Copyright 2010
American Association for the Advancement of Science.)

A single-crystal X-ray diffraction analysis of 26 revealed that absorption spectra of most TTF-annulated porphyrins prepared
the annulated porphyrin adopts a nonplanar saddle-like to date reveal spectral features typical of native porphyrins
conformation in the solid state (Figure 31a,b). The distortion (Figure 32). However, in some cases, a near-IR (NIR) absorption
from planarity was thought to reflect the presence of the bulky is seen that is broad and weak and is thought to originate from a
peripheral substituents126,128−132 (viz., four ethyl groups at β- transfer of charge from the TTF subunit(s) to the central
positions of the pyrroles and four adjacent phenyl groups at the porphyrin. Based on DFT calculations, molecular orbital (MO)
meso positions) and the steric clashes they would impose were diagrams of the singly substituted derivatives 28-H2 and 28-Zn
the macrocycle to adopt a more planar arrangement. were constructed. Analysis of these orbitals revealed that the
The free-base and zinc(II) complexes, 31-H2 and 31-Zn, also HOMO is predominantly concentrated on the TTF unit,
proved amenable to X-ray diffraction structural analysis. These whereas the LUMO is centered primarily on the macrocyclic
analyses revealed only modest deviations from ring planarity as core (Figure 33). This prediction lends credence to the notion
compared to what was seen in the case of 26. Moreover, the that an intramolecular CT phenomenon might be observed upon
meso-pentafluorophenyl groups of 31-H2 and 31-Zn were found photoexcitation. Calculations also revealed that the energetically
to lie almost perpendicular to the mean porphyrin planes (cf. degenerate HOMO − 1 and HOMO − 2 represent what are
Figure 31c−f). The differing structural effects seen for 31-H2 and essentially paired symmetric combinations (a1u and a2u) of the
31-Zn relative to 26 were rationalized in steric terms. porphyrin-centered MOs. The degenerate LUMO and LUMO +
TTF-annulated porphyrins have attracted interest because of 1 were also predicted to be composed primarily from porphyrinic
the possibility that they might display intramolecular donor− MOs with eg symmetry. The observed porphyrin-like optical
acceptor effects in either the ground or excited state. The properties seen for test TTF-porphyrins, such as a strong Soret
2666 DOI: 10.1021/acs.chemrev.6b00375
Chem. Rev. 2017, 117, 2641−2710
Chemical Reviews Review

Figure 25. Schematic representation of photoinduced electron transfer


proposed to occur within supramolecular ensembles formed from the
Figure 23. (a) Near-IR absorption spectral changes that are taken as
electron-rich TTF-calix[4]pyrrole 5 and electron-deficient porphyrin
evidence of ET from 4 (5.0 × 10−5 M) to Li+@C60 (5.0 × 10−5 M) that
carboxylate salts. (a) PET from electron-rich 5 to PA. (b) Dual PET
occurs when THACl is added to a benzonitrile solution of the
from both 5 and PA2 to electron-deficient Li+@C60.
components in question. (b) EPR spectrum (benzonitrile at 298 K) of
the radical ion pair Li+@C60•−⊂4•+·Cl− generated as the result of ET
from 4 (5.0 × 10−5 M) to Li+@C60 (5.0 × 10−5 M) in the presence of
THACl (3.0 × 10−4 M). (c) EPR spectrum of the radical ion pair Li+@
C60•−⊂4•+·Cl− recorded at 77 K. (d) Expanded view of the magnetic
field region highlighted by the red rectangular frame in panel c.
(Reproduced with permission from ref 116. Copyright 2011 American
Chemical Society.)

Figure 24. X-ray structure of the supramolecular complex with a net


stoichiometry of Li+@C60⊂4·Cl−: (a) side view and (b) bottom view.
(Reproduced with permission from ref 116. Copyright 2011 American
Chemical Society.)

band and weaker Q-bands, could thus be rationalized using a


typical Gouterman four-orbital treatment.133,134 Figure 26. (a) Vis−NIR spectral changes consistent with electron
Incremental synthetic addition of TTF units to the porphyrin transfer from 5 (100 mM) to Li+@C60 (100 mM) as seen in the presence
core produces a monotonic red shift in the absorption bands and of increasing concentrations of PA2 in benzonitrile. Inset: Expanded
a decrease in the molar absorption coefficient associated with view of the absorption spectrum over the 600−1600 nm spectral region.
(b) Plots of absorbance at 1035 and 680 nm vs concentration of PA2.
these transitions (see Table 7). These changes were rationalized
(Reproduced with permission from ref 119. Copyright 2015 Royal
in terms of enhanced intramolecular CT effects. The presence of Society of Chemistry.)
four electron-deficient pentafluorophenyl groups (attached to
the meso positions of the porphyrin cores in the cases of 28−31)
serves to enhance this intrinsic CT character (vide infra). In the significantly affect the position or intensity of the intramolecular
presence of electron-rich TTF substituents, the electron- CT (ICT) transitions.127
deficient pentafluorophenyl groups are thought to play an Relative to unsubstituted tetraphenylporphyrin (TPP), an
important role in regulating the observed intramolecular charge- enormous quenching of the fluorescence (greater than ∼90%)
transfer events. On the other hand Zn2+ metalation does not was seen for most test TTF-porphyrins. Such findings are fully
2667 DOI: 10.1021/acs.chemrev.6b00375
Chem. Rev. 2017, 117, 2641−2710
Chemical Reviews Review

Scheme 16. Synthesis of a Schiff-Base TTF-Calix[4]pyrrole consistent with intramolecular electron transfer. When the TTF
Receptor and Its Bimetallic Pd(II) Complex units of 23 and 24 were oxidized by chemical means (e.g., FeCl3),
the emission was (partly) regenerated relative to a comparable
porphyrin control. The excited-state dynamics of the TTF
porphyrins 28−31 were probed using femtosecond transient
absorption (fs-TA) spectroscopy; these studies were carried out
for both the free-base and Zn2+-complexed forms and revealed
that all of the TTF-porphyrins subject to study were
characterized by excited-state lifetimes falling within the
picosecond time scale. These findings were rationalized in
terms of photoinduced charge separation followed by very fast
charge recombination, which serves to generate the ground-state
species. A summary of the photophysical properties of
representative TTF-annulated porphyrins, along with those of
a reference TPP system, is provided in Table 7.
The fusion of TTF units to the periphery of a porphyrin core
can have a significant effect on the driving force for charge-
separation processes. To assess the underlying redox potentials
that were expected to define the thermodynamics of photo-
induced and thermal ET, square-wave voltammetry (SWV)
analyses were used for many TTF-annulated porphyrins.135
Typically, the characteristic redox signals associated with the
production of the TTF radical cation (TTF•+) and dication
(TTF2+) are observed at relatively low potentials (Table 8).
These potentials vary with structure. Moreover, it was found that
the molecular symmetry of the TTF-annulated porphyrins (e.g.,
28−31) has an influence on the electrochemical behavior, as
reflected in the peak patterns observed in the TTF oxidative
Figure 27. Solid-state structure of the Pd(II) complex of a TTF-Schiff- region (cf. Figure 34 and Table 8).
base calix[4]pyrrole (17·Pd2) viewed along the crystallographic (a) a The monosubstituted congeners 28 show two redox waves for
axis and (b) c axis. (This figure was redrawn using data that were oxidation and reduction. In the case of the trans-orientated bis-
originally published in ref 120.)
TTF-porphyrin 26, redox patterns similar to those in 28 were
seen. This congruence led to the suggestion that, in 26, little
Scheme 17. Synthetic Route Leading to a TTF-
intramolecular interaction exists between the two TTF units in
Calix[2]pyrrole[2]thiophene Receptor
the singly oxidized form (i.e., 26•+). In the case of the bis-TTF
porphyrin 29-H2, upon oxidation, complicated oxidative waves
are seen in the SWV results. Although 26 and 29 were not
analyzed under identical conditions, this difference is never-
theless considered to be consistent with the two TTF units
present in 29 being able to communicate electronically through
the intervening porphyrin core.
Similar complexity was seen in the higher derivatives of 30-H2.
A regular positive shift in the reduction signals (Figure 34a) was
also observed in the case of the free-base derivatives as the
number of TTF subunits attached to the porphyrin core
increased.127 These findings were ascribed to an electronic
overlap effect, rather than steric factors.
After metalation, electronic communication between the
separate TTF units becomes restricted, and the individual TTF
sites act as separate redox centers under conditions of oxidation.
Thus, for the zinc(II) complexes, only two signals, corresponding
to TTF•+ and TTF2+, are observed in the oxidative region
(Figure 34b).
Compound 32 is another TTF-annulated porphyrin. It was
synthesized with a view toward exploring whether deaggregation
could be achieved by metal−ligand coordination within a
donor−acceptor-based hybrid supramolecule.136 In dimethyl
sulfoxide (DMSO)/D2O solution, porphyrin 32 exhibits
Figure 28. Solid-state structures of the bis-TTF-calix[2]pyrrole[2]- appreciable aggregation (due to the presence of phenolic −OH
thiophene 18, showing a Pacman-like binding mode with a bound TNB groups at the periphery) and gives rise to spherical particulate
guest: (a) side view and (b) top view. (This figure was redrawn using morphologies, as inferred from scanning electron microscopy
data that were originally published in ref 121.) (SEM) studies (Figure 35). Axial coordination of an imidazole-
functionalized-fullerene (C60Im) guest to the zinc(II) center of
2668 DOI: 10.1021/acs.chemrev.6b00375
Chem. Rev. 2017, 117, 2641−2710
Chemical Reviews Review

Scheme 18. Synthetic Route Used to Obtain a Pacman-Style Calix[2]pyrrole[2]thiophene Receptor System

c,137,138 was reported recently. These systems were obtained in


exceptionally high yield (70−96%) using a procedure that
involved as its key step the nucleophilic substitution of an
aromatic ortho-diamine with a β-diketo-substituted porphyrin (L
or M). It was found that condensation of porphyrin L with
TTFDA gave 33a−c (Scheme 21A), whereas reaction between
porphyrin M and TTFTA, a TTF-tetraamine counterpart,
yielded the TTF-annulated triad 34a−c (Scheme 21B).
As proved true for the directly annulated TTF-porphyrins
discussed above (e.g., 23), the fluorescence intensity and lifetime
of the TTF-pyrazino-porphyrin 33a were decreased relative to
those of a simple porphyrinic control. This reduction was
ascribed to photoinduced electron transfer (PET) from the TTF
unit to the porphyrin core when the system was subjected to
photoexcitation. The fluorescence intensity of 33a increased
upon addition of the chemical oxidant Fe(ClO4)3. Presumably,
this produces an oxidized TTF unit that is a less effective redox
partner for PET. Treatment with a mild reductant restored the
Figure 29. Solid-state structures of the bis-TTF-calix[2]pyrrole[2]- initial, less fluorescent state.
thiophene 19 in its uncomplexed “free” state and after TCNQ binding. Chemical oxidation of 33a with less than 1 equiv of
The electron-deficient TCNQ guest is bound in a Pacman-like binding [Fe(bpy)3](PF6)3 (where bpy = 2,2′-bipyridine) was found to
mode and presumably stabilized by donor−acceptor interactions lead to a concentration-dependent decrease in the intensity of
involving the TTF “arms” of the receptor. Residual solvent and one of the absorption band at ca. 600 nm and the concomitant
the TCNQ molecules are deleted for clarity. Views along the
crystallographic (a) a axis and (b) b axis. (This figure was redrawn
emergence of a new broad absorption band peaking at 830 nm,
using data that were originally published in ref 122.) characteristic of a TTF radical cation (TTF•+) (Figure 38a). The
Soret band was bathochromically shifted while becoming sharper
and more intense. All of these observations led to the suggestion
32 leads to attenuation of the intermolecular aggregation effects that the influence of the porphyrin is minor during the first
(Figure 36). oxidation process, which occurs mainly on the TTF unit. When
Detailed theoretical calculations involving 32 and its supra- 33a was treated with ≥1.4 equiv of [Fe(bpy)3](PF6)3 (Figure
molecular complex 32·C60Im provided insights into their 38b), the TTF radical absorption band decreased, and broad
geometry and electronic nature. As shown in Figure 37a,b, the absorption bands in the 600−1000-nm region emerged. A
HOMO of 32 is mainly localized on the TTF subunits with weak decrease in the intensity of the Soret band was also seen. These
contributions from the porphyrin π-system; however, the changes are attributed to the formation of a porphyrin radical
LUMO is predominately concentrated on the macrocyclic cation. The porphyrin was thus considered to be intimately
core. In complex 32·C60Im, the HOMO is located mainly on involved in the second electrochemical process. Electrochemical
the TTF substituents, whereas the LUMO is all but completely data for representative quinoxaline-fused TTF-porphyrins are
localized on the fullerene group (Figure 37c,d). On this basis, provided in Table 9.
photoinduced ET from the TTF subunits to the C60 moiety Photoinduced electron transfer (PET) can be induced readily
would be expected. within the supramolecular complex formed by treating dyad 33b
Ultrafast intrasupramolecular photoinduced charge separation with a pyridine-appended C60 derivative, C60Py (Figure 39).
between the TTF units and the ligated fullerene was established Initial evidence for the PET process came from state emission
in 32·C60Im by means of emission and transient absorption studies carried out in o-dichlorobenzene, and a gradual
spectroscopic techniques. These studies revealed that the rate quenching of the fluorescence intensity of 33b at 640 nm was
constants for charge separation (kCS) and charge recombination found upon treatment with C60Py (Figure 40). This quenching is
(kCR) were 1.4 × 1011 and 2.5 × 106 s−1, respectively. The lower ascribed to the formation of a singlet excited species, 33b*, that
kCR value was taken as an indication of charge stabilization within then undergoes photoinduced electron transfer to the C60Py
the assembled donor−acceptor conjugate; this was ascribed to an acceptor. A binding constant corresponding to the formation of
electron-transfer−hole-transfer mechanism. the supramolecular complex deemed predicative to PET was
2.4.2. Quinoxaline-Fused TTF-Porphyrins. A series of obtained from a Benesi−Hildebrand plot constructed from the
pyrazine-bridged TTF-annulated porphyrins, 33a−c and 34a− fluorescence titration data. The value derived in this way (7.20 ×
2669 DOI: 10.1021/acs.chemrev.6b00375
Chem. Rev. 2017, 117, 2641−2710
Chemical Reviews Review

Scheme 19. Synthesis of Expanded TTF-Calix[n]pyrroles

Table 6. Effective Association Constants (Ka, M−1)


Corresponding to the Interaction between the BTTF-
Calix[n]pyrroles (Compounds 5, 20, and 21) with
Nitroaromatic (TNB and TNT) Guests and the TBA+ Salts of
Selected Anions
5 20 21
Cl− 520000 13000 5300000
Br− 21000 8300 34000
I− 480 ndb 1800
MeCOO− 140000 22000 870000
H2PO4− 100000 14000 110000
HSO4− 140 2650 7000
TNB 2800, 17000c 870 1200
TNT 570, 2600c 220 480
a
All Ka values were obtained from UV−vis spectroscopic titration
experiments carried out in CHCl3 at 296 K. The receptor/substrate
binding stoichiometries are either 1:1 or 1:2 (when one or two Ka
values are listed), and were determined on the basis of curve fits or
continuous-variation (Job) plots. The anions were studied in the form
of their tetrabutylammonium (TBA+) salts. The estimated errors are
≤15%. bNot determined. cData from ref 95.
Figure 30. Crystal structures of the chloride anion complexes of a set of
congeneric BTTF-calix[n]pyrroles (where n = 4, 5, and 6): (a) 5, (b) 20,
and (c) 21. The tetraalkylammonium counterions are omitted for
104 M−1 for studies in o-dichlorobenzene) proved remarkably clarity. (This figure was redrawn using data that were originally
high. published in ref 123.)
Evidence of a charge-separated state being formed following
photoexcitation of the supramolecular triad 33b·C60Py came 2.4.3. TTF-Annulated Expanded Porphyrins. Expanded
from femtosecond transient absorption spectral studies. TTF-porphyrins, such as the [28]TTF-hexaphyrin, 35, were
Increases in the absorbance intensity at 1000 nm due to the prepared by methods used to prepare conjugated oligopyrroles
C60Py•− radical anion and at 620−700 nm due to the radical lacking annulated TTF subunits (cf. Scheme 22).139 System 35
cation 33b•+ were observed with time. The electron-transfer was found to undergo PET reactions that could be controlled
complex (charge-separated state) was estimated to have a through a unique process involving conformational switching of
relatively long lifetime (385 ps) in toluene. the hexaphyrin core between Hückel figure-of-eight (35-H) and
2670 DOI: 10.1021/acs.chemrev.6b00375
Chem. Rev. 2017, 117, 2641−2710
Chemical Reviews Review

Chart 3. Structural Formulas of Various TTF-Porphyrins and Model Porphyrins

Möbius-twisted (35-M) conformers. This finding provides a The electronic features of the TTF-annulated macrocycles 36
complement to recent studies designed to probe the limits of and 37 could be perturbed by a simple protonation (Figure
Möbius aromaticity, where more classic expanded porphyrins 43a,b). Protonation of the pyrrolic nitrogen centers also served
have played a prominent role.140−142 As a general rule, large to enhance the electron-accepting ability of these macrocycles, an
meso-substituted expanded porphyrins can adopt a variety of effect ascribed to the presence of a positively charged core.
conformations, with several of the well-characterized forms, such Therefore, enhanced intramolecular CT processes were expected
as figure-of-eight and twisted, giving rise to different manifes- in the case of protonated forms. Detailed theoretical calculations
tations of aromaticity. on both the neutral and protonated forms of 36 and 37 revealed
In nonpolar solvents (e.g., CH2Cl2), 35 adopts a figure-of- the HOMOs to be situated on the TTF units (Figure 44a,c,e,g),
eight conformation and displays features consistent with those of whereas the LUMOs are mainly localized on the macrocyclic
a weakly antiaromatic Hückel species, as inferred from spectral core and in part on TTF units (Figure 44b,d,f,h). The
studies and an X-ray crystallographic analysis (Figure 41a,b). In distribution of LUMOs changed substantially in both the cases
contrast, in more polar media (e.g., MeCN) or at low upon protonation (Figure 44f,h). Such findings are consistent
temperature in solvents of intermediate polarity, such as with the notion that thermal charge transfer in 36 and 37 would
tetrahydrofuran (THF), topological switching to a Möbius- take place more efficiently in the presence of TFA. This is also
twisted geometry occurs, as evidenced from the characteristic inferred from the lower-energy absorption bands and active EPR
spectral features143 that mirror those seen in Möbius aromatic signals at about g = 2 seen under ambient conditions in the case of
[28]hexaphyrins (e.g., emergence of an intense and sharp Soret the protonated forms H2·362+ and H2·372+ in CH2Cl2.
band) (Figure 42). Evidence that a charge-separated species was
2.5. ExTTF Porphyrins
formed upon photoexcitation came from an EPR study carried
out at 77 K under conditions of photoillumination. Another approach to combining the chemistry of TTF with that
Calculations revealed that both the HOMO and LUMO of the of porphyrins involves the formal insertion of a porphyrin core
Hückel form, 35-H, are delocalized over the full hexaphyrin π- between the two 1,3-dithiole rings subunits that make up the
circuit, as would be expected for a 28-π-electron species. In TTF molecule (e.g., exTTFs; compounds 38a−d in Scheme
contrast, the HOMOs (up to HOMO − 6) of the twisted form of 24).145 These formal electron-rich quinoidal-porphyrin ana-
35-M are localized on the TTF unit, with the LUMO mainly logues were found to exhibit intriguing redox and photophysical
delocalized over the hexaphyrin core. It is thus believed that this features. For instance, when subjected to two-electron oxidation,
conformer gives rise to more effective PET. the initial quinoid-type porphyrin spacers show features
The core-modified TTF-annulated porphyrin 36 and its 26-π- consistent with aromaticity as inferred from spectroscopic
electron expanded rubyrin analogue 37 have also been studies and the observation that the initial saddle-shaped
reported.144 Compound 36 exhibits features consistent with nonaromatic species become planar as befits an aromatic
intramolecular CT. The syntheses of 36 and 37 are summarized dicationic porphyrinoid structure.
in Scheme 23. The preparations of these two products mirror one A particular advantage of these systems is that their inherent
another, except that different 1:1 mixtures of precursors were properties could readily be controlled by cation complexation.
employed. In contrast to what is true for the figure-of-eight form Scheme 25 illustrates how this coordination chemistry and
of 35 (which contains six fused TTF subunits), expanded subsequent ligand binding to complex 38d can be used to induce
porphyrin 37 contains only two TTF-pyrroles. As a consequence, structural changes that, in turn, modulate the electronic
it was expected to exist as a planar aromatic system. properties of the system as a whole.
2671 DOI: 10.1021/acs.chemrev.6b00375
Chem. Rev. 2017, 117, 2641−2710
Chemical Reviews Review

Scheme 20. Synthetic Schemes Giving Rise to Representative TTF-Porphyrins

When a chloride anion is bound to the Zn(II) center of 38d, transfer, demonstrating the necessity of the Zn−halide
the electron-donor ability of the complex is enhanced. This coordination bond for the forward reaction. It also demonstrates
results in thermal electron transfer from 38d·Cl− to Li+@C60, the reversible nature of the system.
which acts as an electron acceptor. This ET event gives rise to a Single-crystal X-ray diffraction structures of 38a and 38c
stable radical-ion-pair complex under ambient conditions. revealed severe saddle-like distortions, with the dithiole rings
Addition of pyridine to this complex promotes electron back- being bent out of the plane and toward one another in the neutral
2672 DOI: 10.1021/acs.chemrev.6b00375
Chem. Rev. 2017, 117, 2641−2710
Chemical Reviews Review

Figure 31. Crystal structures of (a,b) 26, (c,d) 31-H2, and its Zn2+ complex (e,f) 31-Zn: (a,c,e) top view and (b,d,f) side view. (This figure was redrawn
using data that were originally published in refs 126 and 127.)

Figure 32. UV−vis−NIR spectra of porphyrins 28-H2−31-H2


recorded in dichloromethane. Inset: Region of the CT bands enlarged
by a factor of 15. (Reproduced with permission from ref 127. Copyright
2013 Wiley-VCH Verlag GmbH.)

forms (Figure 45b,d). In contrast, the structure of the two-


electron-oxidized form of 38c (i.e., 38c2+ with triflate as the
counteranion) is planar. These structural changes are consistent
with a system that changes from nonaromatic to aromatic upon
oxidation (Figure 45f). A relatively large two-photon absorption Figure 33. Representative MO diagrams for 28-H2 and its Zn2+ complex
(TPA) cross-section value was found for 38a2+ (1200 GM), 28-Zn, supporting the intramolecular ET phenomena in TTF-
which is much higher than those typically seen for porphyrins porphyrins. (Reproduced with permission from ref 127. Copyright
(<100 GM). This is consistent with the proposed aromatic 2013 Wiley-VCH Verlag GmbH.)
character.
It was proposed that the dicationic species 38d2+ could be a
ments. It was observed that the decay of the charge-separated
component in a three-component supramolecular ensemble that
state produced from this three-component system obeyed first-
would give rise to a long-lived photoinduced charge-separated
order kinetics. An extremely long-lived charge-separated state,
state. The other partners in this system were tetraanionic zinc(II)
with a lifetime of up to 83 ms in benzonitrile at 298 K, was
porphyrin-tetrasulfonates (PZn4−) that would associate through
observed. This long lifetime is thought to reflect the spin-
strong electrostatic interactions to produce a triad of net
forbidden nature of the electron back-transfer process, as well as a
stoichiometry PZn4−·(38d2+)2 in benzonitrile.146 The chemical
small electron-coupling term.
structures of the components along with the supramolecular
complex are shown in Figure 46. An X-ray crystal structure 2.6. Other Porphyrin−TTF-Based Donor−Acceptor Systems
analysis revealed a slipped-sandwich type 1:2 supramolecular 2.6.1. Donor−Acceptor Systems Based on Porphyrins
ensemble wherein the anionic porphyrin PZn4− is sandwiched by with Peripheral TTF Substituents Connected through
two separate dicationic species, 38d2+ (Figure 47). Flexible Organic Spacers. A number of synthetic porphyrin-
When this supramolecular ensemble is subjected to photo- based D−A ensembles based on the use of TTF units have been
excitation, efficient PET from a PZn4− subunit to the triplet developed through meso functionalizations (Chart 4). Meso-
excited state 3[38d2+]* occurs to afford charge-separated triplet TTF functionalization provides an important complement to the
states, as revealed by laser flash photolysis and EPR measure- peripherally annulated TTF-porphyrins in which the TTF units
2673 DOI: 10.1021/acs.chemrev.6b00375
Chem. Rev. 2017, 117, 2641−2710
Chemical Reviews Review

Table 7. Photophysical Data for the TTF-Annulated Porphyrins and a Reference Compound, Tetraphenylporphyrin (TPP)
compd Soret band λabs (nm) Q-band λabs (nm) ε (M−1 cm−1) at Soret band λfl (nm) Φfl τsc (ps)
TPP 420 512, 546, 590, 648 460000 650, 713 0.11 −
23 420 520, 610 186000 650, 710b − −
25 430 520, 590 625000 680, 740b − −
26 441 483, 538, 605 261000 701, 783b − −
27 440 525, 550 696000 710, 770b − −
28-H2 421 511, 542, 587 467000 647, 713 ≤0.01 4, 112, long
28-Zn 427 547, 575 519000 593, 645 ≤0.01 2, 8, long
29-H2 429 515, 547, 575 381000 − − 2.6, 20
29-Zn 432 547, 575 412000 − − 1.5, 7
30-H2 431 519, 554, 606 332000 − − 1.2, 13
30-Zn 438 547, 577 346000 − − 1.2, 6
31-H2 438 526, 560, 606 283000 − − 1, 10
31-Zn 443 547, 581 297000 − − 1, 5
a
Measurements were performed at 298 K in tetrahydrofuran for compounds 23−27 and in CH2Cl2 for compounds 28−31. bEmission maxima [in
(THF), 298 K] of compounds 23−27 were measured after the addition of FeCl3 as a potential oxidant. cTA measurements were carried out in
toluene at 298 K. For compounds 28−30, the excitation wavelength was 510 nm for the free-base forms and 540 nm for the Zn2+ complexes. For
compound 31-H2, excitation was carried out at 560 nm; for 31-Zn. an excitation wavelength of 580 nm was employed.

acceptor in opposing meso positions.147,148 The synthetic route


leading to compounds 40a,b is given in Scheme 26. The
combination of substituents appears to have little effect on the
ground-state optical features of the system. For instance, there is
practically no significant interaction between the TTF unit, the
porphyrin core, and the C60 moiety. This was evidenced by an
absorption spectral analysis of these triads, which revealed an
essentially a linear combination of the spectral features of the
various subunits. In the excited state, the fluorescence of the triad
40a is significantly quenched compared to that of 39. This was
ascribed to photoinduced electron transfer (PET) as seen in
many TTF-porphyrinoid combinations. Presumably, the pres-
ence of a fullerene acceptor (present in 40a but not 39)
contributes further to this effect and to the formation of a charge-
separated state.
The putative singlet excited state of porphyrin of 40a was
probed by time-resolved optical methods. These analyses
revealed spectral features consistent with the formation of an
intermediate charge-separated state, TTF-P•+-C60•−, that decays
with a lifetime of 25 ps. The presence of a zinc(II) center in 40b
affects the ultrafast electron transfer from the singlet porphyrin to
the C60 because of its inherent acceptor property. In both 40a
and 40b, further internal electron-transfer processes yield the
final charge-separated state, TTF•+-P-C60•− that decays with a
Figure 34. Square-wave voltammagrams of selected TTF-porphyrins as lifetime on the order of 660 ns in 2-methyl-THF for both
determined in CH2Cl2 containing 0.1 M tetra-n-butylammonium systems. The generation of relatively longer charge-separated
hexafluorophosphate as the supporting electrolyte: (a) free-base states could reflect a lowest triplet state for the species that is
derivatives 28-H2−31-H2 and (b) their corresponding Zn2+ complexes energetically higher than that of charge-separated species,
28-Zn−31-Zn. (Reproduced with permission from ref 127. Copyright TTF•+-P-C60•−.
2013 Wiley-VCH Verlag GmbH.) Charge separation within 40a was further analyzed by time-
resolved electron paramagnetic resonance (TREPR) spectros-
share the porphyrin plane (vide supra). This is because meso copy.149 In 2-methyl-THF glass (at 10 K) and in a crystalline-
substitution can serve to reduce or sever the electronic phase matrix, the evolution of the spectrum revealed that the final
interactions between the individual D−A components, which charge-separated species, TTF•+-P-C60•−, was formed from the
might be expected to affect, among others, intramolecular singlet TTF-P*-C60 through an intermediate charge-separated
electron-transfer events. In fact, depending on the distance and state, TTF-P•+-C60•−.
orientations of the TTF units connected to the meso positions of Further insights into the photoinduced electron-transfer
the porphyrins, very different excited-state dynamics, redox features of TTF porphyrins came from the study of meso-
behavior, and nonlinear optical properties might be expected. functionalized, TTF-appended-porphyrins (i.e., 41a,b−44)
In pioneering work, Gust and co-workers synthesized several characterized by differing substituents and linkage patterns
molecular systems containing a porphyrin core attached to a (Chart 4). Porphyrin 41a and its Zn(II) complex 41b, bearing
TTF-donor alone (39) or in combination with a C60 (40a,b) TTF subunits attached directly to the meso positions, were
2674 DOI: 10.1021/acs.chemrev.6b00375
Chem. Rev. 2017, 117, 2641−2710
Chemical Reviews Review

Table 8. Redox Formal Potentialsa,b,c (V vs Fc/Fc+) of TTF-Annulated Porphyrins


compd Ered2 (V) Ered1 (V) Eox1 (V) Eox2 (V) Eox3 (V) Eox4 (V) Eox5 (V) ref
PnS-TTF-pyrrolea
− − −0.010 +0.445 − − − 126
PrS-TTF-pyrroleb − − +0.284 +0.661 − − − 127
TPPa −2.010 −1.675 − − − − − 126
23 −1.640 −1.340 +0.275 +0.500 − − − 126
24c −1.650 −1.385 +0.040 +0.405 − − − 125
25 −1.810 −1.570 −0.150 −0.020 +0.360 − − 126
26 −1.770 −1.510 +0.030 +0.270 − − − 126
27 −1.660 −1.320 +0.080 +0.150 +0.250 +0.400 +0.510 126
P-H2 −1.481 −1.055 − − − − − 127
P-Zn −1.528 −1.117 − − − − − 127
28-H2 −1.625 −1.257 +0.124 +0.466 − − − 127
28-Zn −1.638 −1.253 +0.178 +0.434 − − − 127
29-H2 −1.538 −1.205 +0.102 +0.318 +0.441 +0.723 − 127
29-Zn −1.692 −1.235 +0.148 +0.468 − − − 127
30-H2 −1.185 −0.963 −0.123 +0.022 +0.164 +0.258 − 127
30-Zn −1.523 −1.229 +0.174 +0.479 − − − 127
31-H2 −1.015 −0.754 +0.452 +0.618 +0.798 − − 127
31-Zn −1.501 −1.189 +0.223 +0.529 − − − 127
a
Measurements for compounds 23, 25, 26, and 27 were carried out at room temperature in THF containing 0.1 M TBAPF6 as the supporting
electrolyte under an Ar atmosphere and using platinum electrodes as the working and counter electrodes. bMeasurements for the remaining
compounds (viz., P-H2, P-Zn, and 28−31) were carried out in CH2Cl2 containing 0.1 M TBAPF6 as the supporting electrolyte under an Ar
atmosphere and using Pt electrodes as the working and counter electrodes. A silver quasireference electrode was used, and the scan rate was 0.02 V
s−1. cMeasurements for compound 24 were carried out in CH2Cl2 under identical conditions to those described footnote a with a scan rate of 0.1 V
s−1.

Figure 35. SEM image showing the spherical aggregation pattern of 32.
(Reproduced with permission from ref 136. Copyright 2010 Royal
Society of Chemistry.)

Figure 36. Chemical structure of the proposed supramolecular complex


reported by Otsubo and co-workers.150 The synthesis of 41a was
formed by 32 and C60Im.
achieved by means of a P(OMe)3-mediated cross-coupling of
4,5-bis(methylthio)-1,3-dihiol-2-one and the meso-thioxo-1,3-
dithiol-4-yl-substituted porphyrin (N), as summarized in Scheme that oxidation reduces the ability of the TTF subunit to function
27. The key precursor N could be prepared by three independent as an electron donor. Interestingly, however, further addition of
condensation methods as shown in Scheme 27. FeCl3 resulted in the quenching of emission; this was ascribed to
There is apparently no interaction between the TTF unit and porphyrin oxidation rendering that particular component
the porphyrin moiety in these compounds in their ground state; ineffective as a photodonor.
presumably, this is due to the orthogonal meso-linking mode. It Shen and co-workers reported a similar class of TTF−
was also observed that the emission was quenched by 82% in porphyrin conjugates with flexible alkyl linkers (42−44).151,152
toluene, compared to the unsubstituted reference porphyrin, Studies of the absorption spectra revealed prominent features
TPP. As is true for other TTF-porphyrins, this was ascribed to diagnostic of CT interactions. Upon photoexcitation, the
electron transfer from the TTF subunit to the porphyrin emission of compounds 42−44 was quenched by 61%, 51%,
chromophore upon photoexcitation. Chemical oxidation of the and 71%, respectively, compared to the reference TPP; again, this
TTF unit present in 41a by 0.25 equiv of FeCl3 in toluene led to was ascribed to intramolecular photoinduced electron transfer.
enhancement of the porphyrin emission by 12% relative to the Within this series, the degree of fluorescence quenching is largest
TTF-annulated porphyrins. This is presumably due to the fact in the case of triad 44, which bears two TTF units trans to one
2675 DOI: 10.1021/acs.chemrev.6b00375
Chem. Rev. 2017, 117, 2641−2710
Chemical Reviews Review

Figure 37. MO diagrams for (a,b) 32 and (c,d) its supramolecular


complex 32·C60Im. (Reproduced with permission from ref 136.
Copyright 2015 Royal Society of Chemistry.) Figure 38. UV−vis−NIR spectral changes seen for a 7.16 × 10−6 M
CH2Cl2 solution of 33a upon successive addition of aliquots of
[Fe(bpy)3](PF6)3: (a) 0−1 and (b) 1−2.2 equiv of [Fe(bpy)3](PF6)3.
another. Such a finding provides support for the appealing (Reproduced with permission from ref 138. Copyright 2012 Wiley-
conclusion that the extent reflects that CT and other related ET VCH Verlag GmbH & Co.)
effects are directly proportional to the number and placement of
the TTF units attached to the porphyrin core, at least within this the resulting photoelectrochemical cell was inferred based on the
congruent set of derivatives. photoaction spectrum of the self-assembled monolayer (SAM)
Stoddart and co-workers reported a photoactive supra- produced by using 45 in this way. The quantum efficiency of the
molecular machine based on a TTF−porphyrin−C60 triad, 45. system was estimated to be ∼1%, based on the observed
This triad was proposed as a model compound that could help set photocurrent output of 1.3 μA cm−2 when the irradiation source
the stage for the use of supramolecular pseudorotaxanes in was a 26 mW cm−2 laser operating at 413 nm.
energy storage applications (Figure 48).153,154 With this goal in The electrical energy generated by this photoirradiated triad
mine, triad 45, which bears disulfide-based anchoring groups system was exploited to drive a supramolecular machine for
linked to the terminal TTF unit, was self-assembled on a gold dethreading of a [2]pseudorotaxane, a 1,5-bis[(2-
electrode (giving a coverage value of 1.4 nm2 per molecule). hydroxyethoxy)ethoxy]naphthalene (BHEEN) encapsulated
Generation of a photocurrent under conditions of illumination of within a tetracationic cyclobis(paraquat-p-phenylene),

Scheme 21. Synthesis of Quinoxaline-Fused TTF-Porphyrins and Porphyrin Arrays

2676 DOI: 10.1021/acs.chemrev.6b00375


Chem. Rev. 2017, 117, 2641−2710
Chemical Reviews Review

Table 9. Formal Redox Potentials (V vs Fc/Fc+)a of electrons generated during the irradiation [i.e., 3.3 × 10−8 mol/
Quinoxaline-Fused TTF-Porphyrins (4.7 × 10−7 mol) ≈ 7%].
A second-generation redox-/photoactive multicomponent
compound Eox4 (V) Eox3 (V) Eox2 (V) Eox1 (V) Ered1 (V) Ered2 (V)
system was reported by Credi and co-workers.155 This system
33a − +0.85 +0.60 a
+0.15 −1.58 −1.77 marries photoinduced redox machinery with an interlocked
33b − +0.64a +0.48 +0.19 −1.03 −1.76b system (464+) that is based on a [2]rotaxane subunit connected
33cc +1.04 +0.80 +0.50 +0.11 −1.44 −1.77 to a light-harvesting porphyrin donor and a C60 acceptor. This
34a − − +0.75d +0.47a −1.62 −1.79 system contains two donor stations, consisting of a TTF
34b − − +0.56d +0.34a −1.71 −2.06 molecule and a 1,5-dioxynaphthalene (DNP) molecule, built
34c − +0.75a +0.61 +0.44a −1.64 −2.06 into the dumbbell component. This molecular shuttle was thus
a
Two overlapping one-electron redox processes are inferred. expected to afford a light-fueled molecular device because of the
b
Irreversible. cPotentials are reported relative to Ag/AgCl in presence of the electron-accepting CBPQT4+ ring. A CT
CH2Cl2; they can be converted to the Fc/Fc+ scale by subtracting interaction between TTF and the CBPQT4+ ring was seen in
0.51 V from the values referenced to Ag/AgCl. dBroad peak.
the resting state, as inferred from the absorption spectrum. The
stability of this resting state is ascribed to a strong π−π CT
interaction between the CBPQT4+ ring and the electron-rich
TTF unit that serves to enforce a thermodynamic advantage for
the TTF portion of the dumbbell residing near the electron-
deficient CBPQT4+ ring. Upon oxidation of the TTF unit,
electrostatic repulsion between the CBPQT4+ ring and the
charged TTF•+ radical ion shifts the thermodynamics of the
system such that binding to the second electron-donating site
(the DNP “station”) is favored. Thus, photoinduced nanoscale
molecular motions are induced within the [2]rotaxane consisting
of 464+ and the CBPQT4+ ring (Figure 49). Irreversible oxidation
processes ascribed to the TTF units, appearing at ca. +0.51 and
+0.91 V (vs SCE) were observed. These could reflect the
complicated conformational flexibility (e.g., folding), which
induces a partial displacement of the CBPQT4+ ring away from
Figure 39. Schematic representation of the supramolecular triad 33b· the TTF station in 464+.
C60Py formed through the axial coordination of C60Py to 33b and its Movement of the CBPQT4+ ring along the linear portion of
proposed intermolecular charge-transfer pathway. 464+ could also be achieved by reduction of the bipyridinium
moiety at a potential of −0.28 V. The resulting destabilization of
the CT interactions with either the TTF or DNP donor units
facilitates free movement along the thread as inferred from
spectroelectrochemical measurements.
In a strategy that bears some analogy to the alkyne-bridged
porphyrin dimers156−158 and TTF-annulated dumbbell porphyr-
in dimers, 34a−c,137,138 discussed earlier, an A−π−D−π−A-type
TTF-porphyrin dimer, 47, was reported by Ogawa and
Nagatsuka.159 In this system, two porphyrins are bridged by a
diacetylene-TTF unit to give dimer 47. Construct 47 exhibits
features consistent with the presence of effective π-extension, as
evidenced by the large two-photon absorption cross sections
determined by steady-state and time-resolved spectroscopies
(nanosecond open-aperture Z-scan method). The remarkably
large two-photon absorption cross-section values of σ2 = 5900
GM at 760 nm in toluene and 7300 GM in benzonitrile for 47 are
consistent with effective CT and ET within the system. These
Figure 40. Fluorescence spectral changes seen upon the addition of
increasing quantities of C60Py to a solution of 33b in o-dichlorobenzene. values are comparable to those seen for a bisporphyrin directly
(λex = 420 nm.) Inset: Benesi−Hildebrand plot. (Reproduced with connected by a butadiyne bond (σ2 = 7700 GM at 860 nm).160
permission from ref 138. Copyright 2012 Wiley-VCH Verlag GmbH & It was also anticipated that oxidation of the push−pull type of
Co.) TTF-annulated porphyrin species would enhance the static first
hyperpolarizabilities as determined by DFT calculations.161
2.6.2. Donor−Acceptor Systems Consisting of Por-
BHEEN⊂CBPQT4+ (see Figure 48). The electrochemical phyrins with Axially Coordinated TTF Derivatives. Metal-
reduction of cyclobis(paraquat-p-phenylene) (CBPQT4+) weak- loporphyrins are venerable systems that have long attracted
ens the CT interactions, resulting in decomplexation of the attention for use in the construction of efficient CT complexes.
[2]pseudorotaxane, BHEEN⊂CBPQT4+. The reaction could be One approach that has been pursued in the context of this
monitored by measuring the fluorescence intensity of the paradigm involves the preparation of supramolecular D−A
BHEEN moiety at a potential of 0 V. Using this indicator, the multicomponent models that rely on axial coordination to
6.7% increase of the intensity with respected to the initial various porphyrin metal centers. By ligating a TTF donor unit or
intensity was considered to be consistent with the molar ratio of an electron acceptor to a metalloporphyrin, it might be possible
2677 DOI: 10.1021/acs.chemrev.6b00375
Chem. Rev. 2017, 117, 2641−2710
Chemical Reviews Review

Scheme 22. Synthetic Route to Expanded TTF-Porphyrin 35 and Its Switching between Limiting Hückel Figure-of-Eight (35-H)
and Möbius-Twisted (35-M) Conformers

Figure 41. Single-crystal X-ray structure of 35: (a) top and (b) side
views. The thermal ellipsoids are scaled to the 50% probability level.
(This figure was redrawn using data that were originally published in ref
139.)

Figure 42. UV−vis−NIR absorption spectra of 35 recorded in various


to maximize the spatial separation between charges generated by mixtures of CH2Cl2 and MeCN. (Reproduced with permission from ref
photoinduced intracomplex electron transfer. A further 139. Copyright 2013 Royal Society of Chemistry.)
attraction of this approach is that modifications based on
changes in the entities coordinated to the metal centers should be pyridyl-substituted TTF derivative ligated to the metal center
facile. In fact, a number of groups have reported supramolecular (Chart 5).
diads and triads (48−54) based on the use axially coordinated The coordination of the TTF unit to ZnP was analyzed by
TTFs as well as, in certain cases, various ancillary electron- means of a UV−vis absorption spectral titration. Analysis of the
acceptor groups (Chart 5). resulting profile revealed a 1:1 binding constant of 4.71 × 104
One example of a dyad that falls within this general paradigm is M−1. Upon binding of the TTF unit (to produce 48), the
compound 48 reported by Zhu and co-workers.162 This system fluorescence intensity of the porphyrin subunit was quenched
consists of a pentacoordinated zinc(II)porphyrin bearing a relative to that of the starting porphyrin (ZnP). This quenching is
2678 DOI: 10.1021/acs.chemrev.6b00375
Chem. Rev. 2017, 117, 2641−2710
Chemical Reviews Review

Scheme 23. Synthetic Routes to 36 and 37 and Schematic Representations of Their Protonated Forms

played lifetimes of 2.2 and 3.1 ns, respectively.164,165 Upon


complexation of a pyridyl-TTF unit to the zinc(II) center in 49
and 50 (binding constants of Ka = 1.0 × 104 and 1.6 × 104 M−1,
respectively), a hole shift from the oxidized form of ZnP to the
TTF entity occurred following photoexcitation (a driving force
of 0.58 eV was calculated for these systems). Fast electron
transfer from the singlet excited ZnP to the AuP+ moiety
occurred within 70 and 80 ps, respectively, as determined by
transient absorption spectroscopy. Hole shift then occurred to
give the final charge-separated state TTF•+-ZnP-spacer-AuP•−.
The lifetimes of the charge-separated states, 30 and 100 ns for 36
and 37, respectively, were enhanced by 10−50-fold compared to
those of the dyads discussed above (Figure 50a). The
photophysical phenomena associated with this compound are
summarized in Figure 50b.
van der Est and co-workers demonstrated a supramolecular
hole-transfer process in a hypervalent methyl phosphorus(V)
octaethylporphyrin bearing an axially coordinated TTF unit (51)
upon photoexcitation. 1 6 6 In this dyad complex, a
organophosphorus(V) porphyrin was used with the thought
that the high oxidation state of the phosphorus center would
facilitate reduction and help achieve efficient hole transfer
Figure 43. UV−vis−NIR absorption spectra of (a) compounds 36 (Scheme 28).
(black) and H2·362+ (red) and (b) compounds 37 (black) and H2·372+ Coordination of a TTF subunit to the phosphorus center (to
(red) as measured in CH2Cl2. Insets: Magnified NIR region and afford 51) led to significant fluorescence quenching, which was
photographs of solutions of the neutral and diprotonated forms of ascribed to hole transfer from the excited phosphorus porphyrin
compounds (a) 36 and (b) 37 under ambient light. (Reproduced with to the TTF subunit (Figure 51a). Evidence for the formation of a
permission from ref 144. Copyright 2015 Royal Society of Chemistry.) radical pair under these conditions came from time-resolved EPR
(TREPR) spectroscopic analyses (Figure 51b). The observation
ascribed to an electron-transfer process, in which the ZnP of a narrow spectrum with a width of 50 G is consistent with a
macrocycle acts as an electron acceptor and the overall transfer is triplet state consisting of a moderately strongly coupled radical
favored by a driving force of −0.33 eV. However, the degree of pair. Fitting of the decay profiles of the EPR signal allowed a
quenching for the coordinated complex 48 was less than that charge-separated lifetime of ∼300 ns to be calculated.
seen for typical TTF-annulated porphyrins (e.g., 23 and 24), a van der Est and co-workers reported several six-coordinated
result that likely reflects the inherently large separation between D−porphyrin−A supramolecular ensembles based on
the TTF subunits and the ZnP site of complexation. aluminum(III) porphyrins.167,168 Aluminum(III) porphyrins
Lateral π-extended supramolecular triads (e.g., 49 and 50) (AlPs) are attractive because they can accommodate two
constructed through the coordination of a pyridyl-substituted different axially tethered redox-active partners. For instance,
TTF unit to the zinc(II) centers of cationic Zn(II)−Au(III) carboxylic acid derivatives can be used to replace the reactive axial
porphyrin pseudodimers were reported by Odobel and co- hydroxide groups typically found in native AlPs, whereas some
workers.163 The two porphyrins in these pseudodimers were bases, such as pyridines, can coordinate to the opposite face of
bridged by either oligophenylene ethynylene (ZnP-OPE-AuP+) the AlP framework. This can provide facile access to polarized
or bisethynyl quaterthiophene (ZnP-quater-AuP+) spacers. As supramolecular complexes bearing D and A units that are
prepared, these linked porphyrins displayed rapid photoinduced vertically arranged relative to the AlP plane (Chart 5).
electron transfer from the ZnP to the AuP+ unit upon Two triads (52a and 52b), consisting of an AlP unit with an
photoexcitation. The resulting charge-separated systems dis- axially bound naphthalenediimide carboxylate (NDI) electron-
2679 DOI: 10.1021/acs.chemrev.6b00375
Chem. Rev. 2017, 117, 2641−2710
Chemical Reviews Review

Figure 44. (a,c,e,g) HOMO and (b,d,f,h) LUMO diagrams for (a,b) 36, (c,d) 37, (e,f) H2·362+, and (g,h) H2·372+. (Reproduced with permission from
ref 144. Copyright 2015 Royal Society of Chemistry.)

Scheme 24. Synthetic Route to an ExTTF-Porphyrin and Its Scheme 25. Schematic Representation of How Cation
Metal Complexes Complexation and Ligand Binding Can Be Used to Control
the Electronic Properties of MTTFP and Its Thermal ET to
Li+@C60

acceptor unit and a pyridyl-TTF donor unit, were constructed in


this way.167 The binding constant of the axially coordinated
pyridyl-TTF molecule was estimated to be 1.1 × 103 M−1 for a
1:1 receptor/substrate stoichiometry in CH2Cl2. Based on
electrochemical studies, photoinduced electron transfer from van der Est and co-workers reported a series of vertical, linear
an excited AlP* singlet state to the NDI, hole stabilization on the supramolecular triads constructed from AlP, TTF-(Ph)n-Py, and
TTF unit, and an electron shift to NDI were expected to be (Ph)m-C60 moieties (compounds 53a−f in Chart 5), where Py
exothermic. stands for pyridine; Ph stands for phenyl; n = 0, 1; and m = 1, 2,
Experimental evidence was put forward to support the 3.169 The formation of these supramolecular triads was
contention that the excited state of the AlP (i.e., singlet AlP*) confirmed by spectroscopic titrations, with Benesi−Hildebrand
is quenched through electron transfer to the NDI moiety and plot analyses being used to estimate formation constants in the
hole transfer to TTF, giving rise to a TTF•+-AlP-NDI•− radical range of (1.0−1.5) × 103 M−1. Upon addition of TTF-(Ph)n-Py
pair. Spin-polarized transient EPR spectroscopy carried out in (n = 0, 1) to the dyad [AlP-(Ph)m-C60, m = 1−3], supramolecular
the nematic phase of a liquid-crystalline solvent (Figure 52a,b), triads formed, as evidenced by a decrease in fluorescence
4-(n-pentyl)-4′-cyanobiphenyl, revealed polarization patterns intensity. This decrease is consistent with electron transfer from
consistent with charge separation occurring through singlet− TTF to 1AlPor*, a change in the intrinsic fluorescence rates, or
triplet mixing. The lifetime of the resulting TTF•+-AlP-NDI•− some combination thereof (Figure 53a).
charge-separated state was estimated to be 200−350 ns In compounds 53a−f, the acceptor unit is a fullerene (C60),
(depending on the choice of spacer) (Figure 52c).168 In contrast, which is known to have a low reorganization energy. In triads of
in an isotropic solvent, such as CH2Cl2 or benzonitrile, the the general structure 53, excitation of the AlP subunit leads to
lifetime was found to be considerably shorter (<10 ns). stepwise, sequential electron transfer between the TTF unit and
2680 DOI: 10.1021/acs.chemrev.6b00375
Chem. Rev. 2017, 117, 2641−2710
Chemical Reviews Review

Figure 45. Crystal structures of (a,b) 38a, (c,d) 38c, and (e,f) 38c·2CF3SO3−: (a,c,e) front view and (b,d,f) side view. (Reproduced with permission
from ref 145. Copyright 2013 American Chemical Society.)

Figure 47. X-ray crystal structure of the supramolecular complex with a


net stoichiometry of PZn4−·(38d2+)2. (This figure was redrawn using
data that were originally published in ref 146.)

base and cationic gold(III) porphyrin acceptors to give triads 54a


and 54b, respectively (Chart 5). In the case of the free-base triad
54a,170 the energy- and electron-transfer characteristics of the
AlP core could be tuned by changing the solvent polarity or by
adding a TTF donor unit. Specifically, it was found that, In
CH2Cl2 or o-dichlorobenzene, singlet energy transfer occurs
from the excited AlP* singlet state to the free-base porphyrin
Figure 46. Chemical structures of the components used to create a acceptor with a rate constant, kEnT, of 1.78 × 108 s−1 in the
supramolecular triad and a representation of the resulting ensemble.146
absence of an added TTF unit. Upon complexation of the TTF
donor, ultrafast electron transfer from TTF to AlP* occurs
the C60 moiety. This gives rise to a charge-separated radical pair following photoexcitation with rate constants, kET, of (8.33 ×
with a relatively long lifetime (i.e., ∼140 ns).169 The time- 109)−(1.25 × 1010) s−1, depending on the specific system in
resolved EPR spectrum of 53 shows the narrow absorption and question. A subsequent electron shift to the free-base porphyrin
emission band pattern, which is typical of a moderately coupled then occurs to generate the radical ion pair TTF•+-AlP•−-H2P.
radical pair with predominant population in the T0 level at early This species subsequently undergoes an intraensemble electron
times (Figure 53b). At later times, the polarization patterns transfer to generate a spatially well-separated TTF•+-AlP-H2P•−
become weaker and invert as a result of spin-selective radical pair (Figure 54). Time-resolved EPR spectroscopy
recombination from states with singlet character. The weak revealed that this latter radical pair is formed through the triplet
distance dependence between the TTF unit and C60 in the state of the free-base porphyrin and that the exchange coupling of
electron back-transfer was considered indicative of a super- the two spins in 54a is ferromagnetic.171
exchange mechanism. In the case of the triad 54b (cf. Chart 5), having a cationic
Modulation of the excited-state dynamics could be achieved by gold(III) porphyrin (AuP+) as the electron acceptor, spectral
replacing the fullerene acceptor by the electron-deficient free- overlap between the emission and AuP+ absorption is
2681 DOI: 10.1021/acs.chemrev.6b00375
Chem. Rev. 2017, 117, 2641−2710
Chemical Reviews Review

Chart 4. Chemical Structures of Compounds 39−47 Containing Porphyrins Bearing Meso TTF Substituents

observed.172 This facilitates quenching of the fluorescence hydrogen-bonding and electrostatic interactions, are now
through intraensemble electron transfer. Because of the greater known. For instance, Fukuzumi and co-workers reported a
electron-accepting ability of a AuP+ porphyrin complex relative highly distorted saddle-shaped diprotonated porphyrin, O
to the corresponding free base, the lowest excited singlet state of (Figure 55a−c), that forms D−A-type supramolecular ensembles
AlP* was quenched by ultrafast electron transfer to AuP+ with through π−π donor−acceptor interactions with a suitable
halflives of 18.5−61.2 ps, depending on the specific choice of electron-rich guest molecules, such as TTF (Figure 56a).173
spacers that serve to link the components. As in the Whereas many coplanar porphyrins have been used as electron
corresponding free-base system discussed above, intracomplex donors, this highly strained porphyrin (a consequence of the
ET processes ultimately give rise to a charge-separated radical dodecaphenyl substituents) can act as an electron acceptor,
pair. The lifetime of the charge-separated state was estimated to making it attractive for the construction of supramolecular D−A
be 1.4−1.5 ns, with the specific value depending on the number charge-transfer systems.
of phenylene spacers in the compound subject to study (i.e., A single-crystal X-ray diffraction analysis of the complex
choice of donor in 54b). formed between TTF and the dication O revealed that the TTF
2.6.3. Supramolecular Nonbonded Donor−Acceptor is encapsulated within the inner cavity created by the peripheral
Ensembles of Porphyrin and TTFs. A variety of supra- phenyl groups on the porphyrin. This results in the formation of
molecular D−A architectures constructed from TTF subunits TTF-entrapped supramolecular porphyrinic nanochannels
and porphyrins through nonbonding interactions, such as (PNCs) (cf. Figure 56b).
2682 DOI: 10.1021/acs.chemrev.6b00375
Chem. Rev. 2017, 117, 2641−2710
Chemical Reviews Review

Scheme 26. Synthesis of Compounds 40a,b

Scheme 27. Synthetic Route Leading to the Meso-Linked Porphyrins 41a,b

The photoconductivity of a single crystal of TTF-PNC was The formation of a 1:2 hydrogen-bonded supramolecular
studied. A photocurrent of 0.7 nA was recorded at an electrical complex (with a formation constant, K, of 1.9 × 1010 M−2) was
field strength of 3.5 × 104 V cm−1. The photocurrent was found confirmed by spectroscopic measurements. From time-resolved
to be dependent on the axis, providing support for the conclusion measurements, an intrasupramolecular PET reaction from the
that the intermolecular π−π stacked arrangement shown in hydrogen-bonded TTF-carboxylate to the porphyrin O was
Figure 56b is correlated with the greatest conduction pathway. inferred. This gave a charge-separated species at room temper-
The supramolecular TTF-PNC complex was also used to ature. The rate constants corresponding to forward electron
construct photoelectrochemical cells (Figure 57a). A photo- transfer (kET) and backward electron transfer (kBET) were
induced current was generated and the maximum incident- determined to be 7.4 × 1011 and 1.8 × 1011 s−1, respectively, in
photon-to-current efficiency (IPCE) performance of the cell was benzonitrile at room temperature.
determined to be 10.1% at 460 nm when an OTE/SnO2/PCN- The supramolecular electron-transfer behavior of the TTF-
TTF anode was employed (Figure 57b), where OTE represents bridged bis(β-cyclodextrin) derivative (TTFCD) was inves-
optically transparent electrode [e.g., F-doped tin oxide (FTO)]. tigated by Liu and co-workers in the absence and presence of
The excited-state dynamics of hydrogen-bonded supra- porphyrins.175 Cyclodextrins (CDs), cyclic oligosaccharides
molecular PET processes were also probed174 using a linked by α-1,4-glucose bonds, have been used to complex a
combination of TTF carboxylate derivatives as the electron- large number of guests within their central cavities.176
donating guests (along with other putative donor molecules) in Combining a TTF-bridged CD dimer with a porphyrin (P)
conjunction with the diprotonated dodecaphenylporphyrin (O). was found to give rise to a unique noncovalent D−A
2683 DOI: 10.1021/acs.chemrev.6b00375
Chem. Rev. 2017, 117, 2641−2710
Chemical Reviews Review

CD) host under buffered aqueous condition. The binding


constant was estimated to be Ka = 3.50 × 105 M−1 with the
underlying data fitting well to a 1:1 binding profile. This is
consistent with formation of a linear supramolecular assembly.
The porphyrin fluorescence was recovered upon treatment of the
ensemble with H2O2 or oxidation by electrochemical means.
This finding provides support for the proposal that the excited
porphyrin is quenched by intraensemble PET (cf. Figure 58).
A similar type of supramolecular nonbonded complex, namely,
ionic CT dyads formed between a cationic porphyrin and an
anionic TTF derivative, was reported by Dai and co-workers.177
In this case, simple mixing of a N-methylated tetrakis(4-
pyridyl)porphyrin (TMPyP) cation and the TTF-bicarboxylate
(L1) or tetracarboxylate (L2) derivatives afforded the ion-pair
complexes TMPyP/L1 and TMPyP/L2 (Figure 59a,b).
Complexes TMPyP/L1 and TMPyP/L2 were characterized
by broad absorption features in the 750−1200 nm spectral region
that were attributed to anion−cation charge-transfer transitions.
Clearly discernible chemical shifts for the β-pyrrolic protons in
the 1H NMR spectrum, along with an EPR signal at g ≈ 2 and
fluorescent quenching in solution, were seen. This was taken as
evidence of complete or partial CT occurring within these
complexes.
Figure 48. Schematic representation illustrating how the light-driven A photoelectrochecmical cell was prepared using TMPyP and
power supply (i.e., the triad 45) provides the electrical energy needed to L1/L2. A photocurrent response was observed under conditions
dethread a [2]pseudorotaxane ([2]PR) wherein BHEEN is encapsu- of photoillumination (Figure 60). This result, which parallels
lated inside the cavity of the CBPQT4+ cyclophane. The curved arrows those discussed above, was rationalized in terms of an effective
show the vectorial electron transfer from the photoexcited porphyrin photoinduced electron-transfer process.
chromophore to the C60 component. This transfer is followed by a
charge shift to the TTF component. Charge neutralization by the Au 2.7. TTF-Functionalized Porphyrazines
electrode then results in a closed circuit. Electron transfer to the 2.7.1. TTF-Annulated Unsymmetrical and Symmetrical
pseudorotaxane leads to dethreading of the reduced CBPQT•/3+ before
the electron is passed on to the Pt counter electrode. (This figure was
Porphyrazines. Porphyrazines (Pzs), or tetraazaporphyrins, are
redrawn using data that were originally published in ref 154.) analogues of porphyrins, wherein the meso carbon atoms are
replaced by nitrogen atoms.178,179 Porphyrazines have attracted
attention as potential optoelectronic materials because of their
nanoarchitecture in aqueous media. The basic recognition desirable photophysical properties, including intense π → π*
chemistry is shown schematically in Figure 58. transitions in the visible region. When TTFs are annulated or
A quenching of the porphyrin fluorescence was observed when otherwise synthetically appended to these macrocyclic systems,
the water-soluble meso-tetrakis(4-sulfonatephenyl)porphyrin interesting hybrid molecular motifs are generated that, in many
(P) was allowed to interaction with the TTF-bridged bis(β- instances, display unique optical and magnetic properties.

Figure 49. Schematic representation of a bistable mechanical switch composed of 46 (in different charge states), CBPQT4+, DNP, and TTF•+.

2684 DOI: 10.1021/acs.chemrev.6b00375


Chem. Rev. 2017, 117, 2641−2710
Chemical Reviews Review

Chart 5. Chemical Structures of Some Supramolecular Di- and Triads Containing Porphyrins with Axially Coordinated TTF
Derivatives

Yin and co-workers reported a series of symmetrical180−182


and unsymmetrical183 TTF-annulated metalloporphyrazines
with various peripheral substitutuents (structures 55−57 in
Chart 6). These compounds were prepared through the
cyclotetramerization of appropriately chosen dicyano-TTF
derivatives. Their optical, electrochemical, and aggregation
properties were then analyzed.
The syntheses of representative unsymmetrical and sym-
metrical porphyrazines are summarized in Scheme 29. The key
preparative step involves the Mg(II)-templated macrocyclization
of a 2,3-dicyano-TTF precursor (e.g., DCTTF1 to synthesize
55b) in an alcoholic solvent at reflux.
The 1H NMR spectra of 56a−c recorded in CDCl3 revealed
broad signals. This broadening was thought to reflect the
presence of a radical contaminant, such as TTF•+, or slow
molecular tumbling on the NMR time scale due to aggregation of
the compound. Broadening of both the Soret and Q-bands was
also noted in the UV−vis spectra; this was attributed to n → π*
electronic transitions involving nonbonding electrons associated
with the peripheral S and N atoms, as well as strong aggregation
resulting from Pz−Pz, Pz−TTF, and TTF−TTF π−π
Figure 50. (a) Nanosecond transient absorption spectra and best fits for
the bleaching of the AuP Soret band at 418 nm for dyad 49 in
interactions. Dynamic light scattering (DLS) measurements
dichloromethane (red); dyad 50 in toluene (blue); and for reference, a provided support for the conclusion that the monomeric forms of
quaterthiophene dyad without a secondary donor (black). All traces these compounds self-assemble in solution to form near-uniform
were obtained after excitation of the Q-band of the ZnP moiety. The aggregates that are approximately 200 nm in size. Further
solutions were deaerated by bubbling with Ar. (b) Summary of the evidence for aggregation came from transmission electron
photophysical processes occurring in the triads produced by complex- microscopy (TEM) experiments.
ation of a TTF subunit. (Reproduced with permission from ref 163. Incorporation of electron-withdrawing ester groups at the
Copyright 2010 Royal Society of Chemistry.) periphery of the TTF units served to reduce the donor properties
of these compounds. This meant that effective ET was seen only
2685 DOI: 10.1021/acs.chemrev.6b00375
Chem. Rev. 2017, 117, 2641−2710
Chemical Reviews Review

Scheme 28. Synthesis of Compound 51

in the case of strong acceptors. For example, when compound


57c was mixed with TCNQ, no CT band was observed in the
600−1000-nm spectral region. However, when a stronger
electron-withdrawing acceptor, namely, 2,3,5,6-tetrafluoro-
7,7,8,8-tetracyanoquinodimethane (F4TCNQ), was used in
conjunction with 57c, CT transitions were observed at 756
and 863 nm (Figure 61a). The formation of a 57c•+/F4TCNQ•−
radical ion pair was confirmed by EPR spectroscopy (Figure
61b).
Electrochemical analyses of compounds 56 and 57 in a
CH2Cl2/MeCN mixture revealed one reduction signal for each
compound. These redox events were ascribed to porphyrazine-
centered processes. In contrast, multiple signals were observed in
the oxidative region, a finding considered to reflect redox events
involving the various subunits present in the overall framework.
Su and co-workers carried out an in-depth theoretical
calculation184 on a simplified form of the TTF-annulated
Zn(II)-porphyrazine ZnTTFPz (56c). This model system was
identical to 56c but lacked the peripheral butyl substituents (so
as to reduce the required computation time). Also examined
were two controls, namely, an unsubstituted normal Zn(II)-
porphyrazine and Zn(II)-TPP. A goal of this study was to
understand the origin of the unusually broad, intense, and red-
shifted Q-band.
The vertical excitation energies of the compounds were
Figure 51. (a) Emission spectra of 51, its oxidized form 51•+, and the calculated using time-dependent density functional theory
reference POEP-OMe (λex = 418 nm). (b) Time-resolved EPR spectra (TDDFT) and compared to the experimental UV−vis
of the reference compound POEP-OMe and 51, recorded at 5 °C in a absorption spectra. On this basis, it was concluded that the Q-
liquid-crystalline matrix consisting of 4-pentyl-4′-cyanobiphenyl (5CB). band of ZnTTFPz is dominated by a TTF → Pz charge-transfer
(This figure was redrawn from data originally published in ref 166.) transition that mixes with those arising from the porphyrazine
2686 DOI: 10.1021/acs.chemrev.6b00375
Chem. Rev. 2017, 117, 2641−2710
Chemical Reviews Review

Figure 52. Room-temperature (300 K) transient EPR spectra of


compound 52 recorded in the nematic phase of a liquid crystal, 4-(n-
pentyl)-4′-cyanobiphenyl. (a) Wide scan showing a broad, very weak
contribution from 3AlP and intense narrow peaks from TTF•+-AlP-
NDI•− 300 ns after the laser flash. (b) Features ascribed to TTF•+-AlP-
NDI•− viewed using an expanded scale: solid line, 300 ns after the laser
flash; dashed line, 1.1 μs after the laser flash. The arrow under the spectra Figure 53. (a) Fluorescence titrations of AlP-(Ph)2-C60 with TTF-Py in
indicates the field position used to record the spectroscopic transients o-dichlorobenzene. Excitation (λex) was effected at a wavelength
used for the lifetime analyses. (c) Signal decay associated with the corresponding to the isosbestic point, 555 nm, as determined by
proposed charge recombination that occurs in complex 52 following independent UV−vis titrations. (b) Room-temperature spin-polarized
photoirradiation. (Reproduced with permission from ref 167. Copyright transient EPR spectra of TTF-(Ph)n-Py-AlP-(Ph)m-C60 triad. The
2013 Wiley-VCH Verlag GmbH.) spectra on the right and left have been extracted from the time and field
data sets associated with the 75- and 550-ns time windows, respectively.
The black traces are experimental spectra, and the red traces are
core. The Q-band of ZnTTFPz was found to mix with other
simulations. (Reproduced with permission from ref 169. Copyright
configurations, meaning that the system could not be analyzed in 2016 Royal Society of Chemistry.)
terms of Gouterman’s classic four-orbital model that has been
used so successfully for the spectral interpretation of many
tetrapyrrolic systems. On the other hand, the match with steady-state absorption spectrum (Figure 62a). These two newly
experiment (maximum error in the excitation energies of 0.17 generated bands correspond to the SOMO → LUMO electronic
eV) provided support for the notion that TDDFT/statistical transition that produces 58•+. Support for this conclusion came
average of orbital potentials (SAOP) performs well in predicting from EPR spectroscopic studies (Figure 62b). Specifically, peaks
the Q- and B-bands, at least of ZnTTFPz. at g = 2.0078 and g = 2.00176 are seen in the EPR spectrum
2.7.2. Porphyrazines Peripherally Modified with TTFs recorded at room temperature in CH2Cl2 that are ascribed to the
through Saturated Spacers. TTF-Thia-Crown-Ether-Func- TTF radical cation (TTF•+)186,187 and the F4TCNQ radical
tionalized Porphyrazine. Yin and co-workers also reported a anion (F4TCNQ•−)188,189 generated as the result of the
TTF-thiacrown-ether-annulated porphyrazine185 (58). This proposed CT process.
compound was obtained using the same basic Mg2+-templated Broadening of the Q-band as well as the B-band was seen in the
macrocyclization process as described above; in this case, case of 58. This broadening was thought to be due to n → π*
precursor Q was employed (cf. Scheme 30). electronic transitions involving the nonbonding electrons of the
The porphyrazine system 58 was investigated by electro- peripheral S and N atoms. An electrochemical analysis of 58
chemical methods and CT studies. As is true for other revealed two one-electron reduction processes at −1.177 and
porphyrazines described earlier in this review, no evidence of −0.937 V that are associated with reduction of the magnesium-
electron transfer from 58 to TCNQ was seen, even when the (II)-porphyrazine core (Table 10). In addition, three signals
latter putative acceptor was present in excess. However, when were seen in the oxidative region. These are ascribed to multiple
F4TCNQ was used, an intermolecular CT complex, assigned to oxidation processes and are thought to reflect interactions
the 58•+/F4TCNQ•− radical ion pair, formed, as inferred from between two neighboring TTF units, which allow species such as
the appearance of two CT bands (λmax = 759 and 863 nm) in the TTF•+, [(TTF)2] •+, and TTF2+ to be stabilized.
2687 DOI: 10.1021/acs.chemrev.6b00375
Chem. Rev. 2017, 117, 2641−2710
Chemical Reviews Review

CHCl3 and MeCN (3:1, v/v) using an excess of TCNQ. Under


these conditions, two new absorption bands at 747 and 847 nm
were observed, in analogy to what was found for 57c (Figure 61a)
and 58 (Figure 62a). These new bands were assigned to
SOMO−LUMO transitions of the TTF cation radical species
present in the radical ion pair produced under these conditions,
namely, 59•+/TCNQ•−.
Electrochemical investigations revealed a total of four redox
waves. One, at −0.693 V vs SCE, was assigned to Pz-core
reduction (i.e., Pz2−/Pz3−). The other three, at +0.563, +0.938,
and +1.218 V vs SCE, were assigned to TTF•+/TTF, TTF+2/
TTF•+, and Pz−/Pz2− oxidative processes, respectively.
2.7.3. Pyrazinoporphyrazine with Peripheral TTF Units
Tethered by Flexible Spacers. A unique type of pyrazino-
porphyrazine derivative (60a−c) containing eight TTF units
attached at the periphery of the macrocyclic core was reported by
Howard and co-workers.191 This compound was synthesized by
subjecting a TTF-tethered 2,3-dicyanopyrazine-based precursor,
S, to cyclotrimerization (Scheme 32).
Figure 54. Proposed energy-/electron-transfer pathway operative in
dyad (AlP-H2P) and triad 54a.
The main precursor, S, was obtained in a straightforward
manner using procedures analogous to those used to prepare
other unsymmetrical TTF derivatives. It was then subjected to
lithium pentoxide-promoted macrocyclization at high temper-
ature. The 1H NMR spectra of 60a−c, recorded in a wide range
of solvents at 20 °C, were characterized by broad signals. The
UV−vis absorption spectra of 60b,c in CH2Cl2 showed a single
Q-band-like absorption feature centered at λmax values of 659 and
664 nm in the cases of 60b and 60c, respectively. Upon addition
of excess iodine to a CH2Cl2 solution of 60b, a new broad low-
energy absorption band emerged beyond 800 nm, a finding that
was interpreted in terms of the formation of TTF•+. The
nonemissive nature of these compounds was explained in terms
of intramolecular CT between the TTF units and the excited
state of the pyrazinoporphyrazine core, in analogy to other TTF-
annulated macrocyclic systems discussed in this review (vide
supra).
Electrochemical analysis of 60a−c in a range of pure and mixed
solvents provided support for the notion that the adjacent TTF
units present at the periphery of the macrocyclic core do not
Figure 55. (a) Chemical structure of O. (b,c) Single-crystal X-ray crystal interact with each other. As a consequence, two clear sets of
structure of O: (b) top view and (c) side view showing the saddle-like oxidation waves are seen that are similar to those of the precursor
distortion from planarity. (This figure was redrawn using data that were S. The reversibility of the oxidation procedure decreases as the
originally published in ref 173.) solubility of the compound in question (i.e., 60a > 60b > 60c)
decreases.
Porphyrazine Peripherally Attached to TTFs through 2.8. TTF-Functionalized Phthalocyanines
Saturated Thia-Ether Bridges. Yin and co-workers reported190
a metalloporphyrazine, 59, wherein eight peripheral TTF units 2.8.1. TTF-Annulated Symmetrical Phthalocyanines.
are connected to the macrocyclic core through saturated thia- Phthalocyanines192−194 (Pcs) are venerable pigments that, like
ether linkages at all of the β-positions of pyrroles. Compound 59 porphyrazines, are structurally related to the porphyrins. The
was synthesized by the magnesium propoxide-promoted cyclo- parent Pc framework contains four isoindole subunits linked by
tetramerization of precursor R in propanol at reflux, as shown in four meso nitrogen atoms. Pcs can be viewed as 18 π-electron
Scheme 31. aromatic systems that act as planar dianionic N4 macrocyclic
The UV−vis spectrum of 59, recorded in a mixture of CHCl3 ligands for a variety of metal cations. In fact, most Pcs are far
and MeCN, was characterized by a broad Q-band centered at 678 more stable in their metalated forms. The meso nitrogen atoms
nm (ε = 58000 M−1 cm−1), as well as a Soret-like transition that alter the MO interactions relative to what is typical for
was thought to reflect transitions involving both the Mg(II)-Pz porphyrins, especially in terms of destabilizing the HOMOs.
ring and the TTF substituents. Broadening of both the Q- and B- One consequence of this modulation is intense Q-band(s) for
bands was attributed to n → π* transitions involving the Pcs that are observed at about 700 nm in the visible region. This
nonbonding electrons of the peripheral S and N atoms that transition has made Pcs attractive for their photophysical
overlap with the pure Mg(II)-Pz ring-derived bands. To properties and made them widely used as functional dyes and
investigate the ability of compound 59 to act as an electron industrial pigments. From an electrochemical perspective, Pcs
donor, a doping experiment was carried out in a mixture of behave as the electron acceptors, particularly when they are
2688 DOI: 10.1021/acs.chemrev.6b00375
Chem. Rev. 2017, 117, 2641−2710
Chemical Reviews Review

Figure 56. (a) Schematic diagram of the supramolecular D−A ensemble that results from treating the diprotonated saddle-shaped porphyrin host O
with the electron donor TTF. (b) Crystal packing diagram of the ensemble illustrating the columnar structure that persists along the crystallographic c
axis. Solvent molecules have been omitted for clarity. (This figure was redrawn using data that were originally published in ref 173.)

Figure 57. (a) Sample arrangements used in photoconductivity


measurements of a single crystal of PNC-TTF. (b) Schematic
description of a photoelectrochemical cell employing an OTE/SnO2/
PNC-TTF anode. (Reproduced with permission from ref 173.
Figure 59. Schematic representation of two different supramolecular
Copyright 2008 American Chemical Society.)
structures formed when cationic TMPyP is exposed to anionic TTF
carboxylates (L1 and L2). (This figure was redrawn using data that were
combined with TTF electron donors. This has made Pcs originally published in ref 177.)
functionalized with TTF subunits desirable synthetic targets.

Figure 58. Schematic representation of the supramolecular ensembles formed from TTFCD and porphyrin P in aqueous media. (This figure was
redrawn using data that were originally published in ref 175.)

2689 DOI: 10.1021/acs.chemrev.6b00375


Chem. Rev. 2017, 117, 2641−2710
Chemical Reviews Review

The absorption spectra of 61d, recorded in THF in both its


neutral form and its tetraradical cationic state, are shown in
Figure 63a. It was demonstrated that the relatively broad Q-band
observed in the neutral form of 61d is due to aggregation
occurring in THF; in contrast, the sharper Q-band centered at
700 nm was ascribed to the monomeric species. In the case of the
symmetrical TTF-annulated Pcs, the emission is quenched in the
ground state. This quenching presumably reflects an efficient
intramolecular reductive electron transfer from the TTF subunits
to the Pc core. The emission intensity increases upon chemical
oxidation by FeCl3. Thus, these systems behave as efficient redox
Figure 60. Energy diagram for components [TMPyP-(L1)4]n and switches. On the other hand, the neutral species 61a−d aggregate
TMPyP used to construct a photocell on an ITO glass electrode. (b) rapidly in solution, which results in a quenching of the typical
Photocurrent responses of crystals of [TMPyP-(L1)4]n and [TMPyP-
(L2)2]n recorded in the presence of 0.1 M Na2SO4 aqueous solution.
phthalocyanine-based emissions at 710 and 786 nm. This aspect
(Reproduced with permission from ref 177. Copyright 2016 Royal of their chemistry might reduce the utility of systems such as 61
Society of Chemistry.) as molecular switches.
Electrochemical studies carried out by means of thin-layer
Chart 6. Chemical Structures of TTF-Annulated cyclic voltammetry revealed that all four TTF units in
Porphyrazines compounds 61a−d are oxidized in two quasireversible steps
involving two four-electron oxidation processes. Within this
series, only the free-base form of 61a shows a reversible one-
electron reduction signal at −0.98 V vs Fc0/Fc+. This signal was
ascribed to reduction of the Pc core (Figure 63b). All other TTF-
annulated Pc derivatives are free of features under analogous
reduction conditions (see Table 11 for details).
After the initial reports by Decurtins and co-workers,195,196
Kimura and co-workers reported a series of symmetrical197−200
TTF-annulated Pcs containing soluble alkyl chains at the α-
positions (cf. compounds 63a and 63b in Chart 7, for example)
and used these compounds to investigate various substituent
effects. The synthesis of these species is shown in Scheme 33B.
Absorption spectral studies revealed a broad CT absorption band
that bears analogy to what is seen for various TTF-annulated
porphyrins (vide supra). This band is thought to originate from
partial electron transfer from the TTF subunits (donors) to the
Pc core (acceptor). Upon addition of iodine to these TTF-
annulated Pcs, the Q-band becomes significantly broadened and
undergoes a prominent red shift (Figure 63c). These spectral
changes were ascribed to the formation of an electron-transfer
complex displaying radical features.
Wang et al. reported201 another type of symmetrical thia-
crown-ether-TTF-functionalized Pc derivative. This compound,
64, was synthesized through the cyclotetramerization of the
corresponding thia-crown-ether-based phthalonitrile precursor
T (Scheme 33C). The absorption spectrum of 64 is characterized
by a TTF-based absorption band that overlaps with a broad Soret
band, a finding that was considered indicative of strong
aggregation in solution. Upon addition of TCNQ to this
compound, a sharp band appears at 400 nm, whereas the Q-band
remains intact (Figure 64a).
Syntheses of phthalocyanine derivatives are often accom- Electrochemical investigations of 64 revealed two oxidation
plished through the self-condensation of phthalic acids, processes that do not apparently involve oxidation of the Pc core;
phthalonitriles, or diiminoisoindole derivatives. These strategies rather, the electrochemical processes in question appear to reflect
have been used to prepare TTF-Pcs. For instance, Decurtins and two separate four-electron oxidation processes associated with
co-workers reported195,196 a series of symmetrical TTF- the TTF units (Figure 64b and Table 11). A slight separation was
annulated Pcs along with their corresponding metal complexes seen for the first oxidation wave that was thought to reflect an
(61a−d in Chart 7). These systems were prepared by interaction between the TTF•+ radical produced by oxidation
macrocyclization of appropriately chosen TTF-annulated Pc with neighboring TTF units, thus generating a species best
precursors in basic media (Scheme 33A). In this study, the represented as [(TTF)2] •+.
authors also demonstrated that those systems that bore long and To gain insight into the sensing capability for alkaline ions,
flexible aliphatic chains exhibited discotic liquid-crystalline electrochemical analyses of 64 were carried out in the presence of
behavior; presumably, this reflects a cofacially stacked arrange- soluble (perchlorate) Na+, Li+, K+, and Cs+ cation salts. These
ment of the extended and planar TTF-Pc cores. studies provided evidence of selectivity for the Na+ ion in that a
2690 DOI: 10.1021/acs.chemrev.6b00375
Chem. Rev. 2017, 117, 2641−2710
Chemical Reviews Review

Table 10. Electrochemical Data of Various TTF-Annulated Porphyrazines


compd Ered3 (V) Ered2 (V) Ered1 (V) Eox1 (V) Eox2 (V) Eox3 (V) Eox4 (V) ref
55a −1.760 −0.980 −0.490 +0.570 +0.910 +1.290 +1.630 183
55b − −1.190 −0.870 +0.440 +0.840 +1.130 +1.250 183
55c −1.790 −1.190 −0.830 +0.480 +0.840 +1.290 +1.860 183
56a − − − − − − − −
56ba − − −0.580 +0.670 +1.040 − − 180
56c − − − − − − − −
57ab − − −0.836 +0.998 +1.313 +1.784 − 182
57bb − − −0.802 +0.972 +1.304 +1.784 − 182
57cb − − −0.813 +0.914 +1.308 +1.786 − 182
58c − −1.177 −0.937 +0.544 +0.876 +1.189 − 185
59d − − −0.693 +0.563 +0.938 +1.218 − 190
a
Measurements were carried out in CH2Cl2/MeCN (5:1, v/v) containing 0.1 M TBAPF6 as the supporting electrolyte under an Ar atmosphere.
Platinum electrodes were used for the working and counter electrodes, and the reference electrode was Ag/AgCl. The scan rate was 100 mV s−1.
b
Measurements were carried out in CH2Cl2/MeCN (4:1, v/v) containing 0.1 M TBAPF6 as the supporting electrolyte under an Ar atmosphere.
Counter and working electrodes were made of Pt and glassy carbon, respectively; a calomel electrode (SCE) was used as the reference electrode. The
scan rate was 100 mV s−1. cMeasurements were carried out in CH2Cl2/MeCN (9:1, v/v) containing 0.1 M TBAPF6 as the supporting electrolyte
under an Ar atmosphere. Counter and working electrodes were made of Pt and glassy carbon, respectively; a calomel electrode (SCE) was used as
the reference electrode. The scan rate was 100 mV s−1. dMeasurements were carried out in CHCl3/MeCN (3:1, v/v) containing 0.1 M TBAPF6 as
the supporting electrolyte. The scan rate was 100 mV s−1. Potentials are reported against SCE.

Scheme 29. Synthetic Routes of Various Unsymmetrical and Symmetrical TTF-Annulated Porphyrazines

large positive shift in redox potential (ΔE1/21 = 80 mV) was seen complexing ability of the crown ether moieties to produce what
for Na+ by differential pulse voltammetry (DPV), whereas no might emerge as a practical, electrochemical-based Na+ cation
appreciable shifts were seen for the other alkali-metal ions tested sensor.
(e.g., Li+, K+, and Cs+). In this particular supramolecular system, 2.8.2. TTF-Crown-Ether-Functionalized Phthalocya-
the redox properties of the TTF unit are combined with the nines. Nolte and co-workers reported classic thia-oxa-crown-
2691 DOI: 10.1021/acs.chemrev.6b00375
Chem. Rev. 2017, 117, 2641−2710
Chemical Reviews Review

Figure 61. (a) UV−vis−NIR spectral changes observed for a 2 × 10−5 M


CH2Cl2 solution of 57c upon the addition of 4 equiv of F4TCNQ. (b)
EPR spectrum of 57c recorded after mixing with F4TCNQ in CH2Cl2 at
room temperature. (Redrawn from data that were originally published in Figure 62. (a) UV−vis−NIR spectral changes seen for a 1 × 10−5 M
ref 182.) CH2Cl2 solution of 58 upon the addition of 4 equiv F4TCNQ. (b) EPR
spectrum of 58 recorded in the presence of F4TCNQ in CH2Cl2 at room
ether-fused metallophthalocyanines, as embodied in the temperature. (This figure was redrawn from data that were originally
symmetric systems 65a and 65b.202,203 These systems were published in ref 185.)
found to self-assemble into helical tapes with nanometer widths
and lengths on the order of a few micrometers. These constructs aggregation under these conditions (Figure 65a). Transmission
were found to display electron conductivity, ion-transport electron microscopy (TEM) revealed the formation of fibers of
features, and liquid crystallinity. The free-base form was up to several micrometers in length, which corresponds to stacks
synthesized by tetracyclization of a TTF-thia-oxa-crown-ether- of ca. 105 molecules (Figure 65b,c). A thorough inspection of the
annulated phthalonitrile precursor, U, whereas its Cu2+ complex, fibers revealed the presence of a very thin bilayer-type motifs
65b, was synthesized from a diiminoisoindoline derivative of U (Figure 65b). These bilayers have widths on the order of 20 nm,
(i.e., V) using CuCl2 as a copper(II) cation source (Scheme 34). which corresponds to approximately 5 times the calculated width
Absorption spectral analyses of 65a were carried out in mixed of the individual molecules. From these observations and initial
organic media consisting of various ratios of CHCl3 and MeOH. modeling experiments, it was suggested that an unusual
These studies revealed that, as the concentration of MeOH was combination of multiple TTF−TTF and TTF−Pc interactions
increased, the Q-bands at 663 and 700 nm underwent a (as opposed to Pc−Pc interactions) underlie the supramolecular
hypsochromic shift eventually giving rise to a broad feature self-association of 65a that leads to the formation of the higher-
centered at 630 nm at a 1:4 (v/v) ratio of CHCl3 to MeOH. This order self-assembled scrolled molecular architectures. Both left
spectral feature was considered indicative of 65a undergoing and right helices are observed (Figure 65c), as would be expected

Scheme 30. Synthetic Route Used to Prepare a TTF-Thia-Crown-Ether-Functionalized Porphyrazine

2692 DOI: 10.1021/acs.chemrev.6b00375


Chem. Rev. 2017, 117, 2641−2710
Chemical Reviews Review

Scheme 31. Synthesis of Compound 59

Scheme 32. Synthetic Route Leading to TTF-Appended Pyrazinoporphyrazines

in the absence of any chiral bias associated with the aggregation identical absorption behaviors irrespective of the number of
process. TTF units attached to the central macrocyclic core. This finding
2.8.3. TTF-Annulated Unsymmetrical Phthalocyanines. was rationalized in terms of π-electronic delocalization within the
Kimura and co-workers reported a series of unsymmetrical TTF- neutral forms of these systems that is not strongly affected by the
annulated Pcs199,204−207 (compounds 66−70; Chart 8) that were outer functional groups (i.e., annulated to the isoindoline phenyl
obtained from a mixture of two nonidentical phthalonitrile groups). The addition of trifluoroacetic acid (TFA) to the TTF-
derivatives. Photophysical, electrochemical, and surface chemical phthalocyanine 66a generated a cationic species and gave rise to a
studies were carried out to characterize these systems. The new band centered at λmax = 740 nm (Figure 66) in the UV−vis−
general synthetic routes used to prepare the unsymmetrical Pcs NIR absorption spectrum.
66−70 are summarized in Scheme 35. Kimura and co-workers also demonstrated that unsymmetrical
Steady-state absorption spectral studies revealed that all of the TTF-fused Pcs, such as 70, would aggregate and that the alkyl
asymmetrically substituted Pcs 66−70 give rise to almost chains would pack closely within Langmuir−Blodgett (LB)
2693 DOI: 10.1021/acs.chemrev.6b00375
Chem. Rev. 2017, 117, 2641−2710
Chemical Reviews Review

Chart 7. Chemical Structures of Representative Symmetrical TTF-Annulated Phthalocyanines

films.204 In these films, the Pc ring planes are oriented on the fan-type birefringence texture observed under conditions
perpendicular to the water surface, whereas the long peripheral of polarized light microscopy. In contrast, incorporation of two
alkyl chains extend into the air. The LB films constructed by 70 TTF units (as in 72) suppresses the mesophase formation. The
exhibit low conductivities in the range of 10−10−10−9 S cm−1 at temperature range of the mesophase of 71, ca. 24 °C, is
room temperature. Upon exposure of the LB films to I2 vapor, significantly less than that of its precursor phthalocyanine (ca. 90
their conductivities increased to 10−5−10−4 S cm−1, an effect °C). Absorption spectroscopic analysis of spin-coated films of
ascribed to the partial oxidation of the monomers within the LB these compounds revealed that their molecular architectures can
films. be modified by oxidation with iodine vapor or by heating.
2.8.4. Phthalocyanines Containing Peripheral TTF Another class of metal-free phthalocyanine derivatives was
Units Connected through Saturated Spacers. There are reported by Bryce and co-workers.209,210 The Pc systems in
numerous reports of Pc derivatives (Chart 9) wherein TTF units question are symmetrically functionalized with eight (e.g., 73a,b)
are attached to the periphery of the macrocyclic core, using either or four (i.e., 74) peripheral TTF units. It was envisaged that the
flexible saturated aliphatic linkers (e.g., 71−75)208−214 or axial presence of saturated alkyl-ether bridges would serve the dual
coordination to the Pc metal centers (e.g., 76).215 purpose of conferring some solubility on the system while
In general terms, these TTF-functionalized Pcs can be allowing the peripheral TTF units the conformational freedom
considered as D−σ−A systems. They were thus expected to needed to achieve coplanarity with the Pc ring and form
act as a new generation of functional materials with interesting intermolecular face-to-face stacks.
solution-state and solid-state properties. They are also of interest The synthesis of compounds 73a,b is shown in Scheme 36A. It
in terms of understanding the importance of various supra- relies on a conventional high-temperature lithium pentoxide-
molecular interactions wherein redox-active groups are attached mediated macrocyclization reaction and involves the use of a
to the highly conjugated planar Pc core in a dendritic fashion. doubly TTF-substituted phthalonitrile synthon (DTTFPCN).
Cook et al. reported two unsymmetrically substituted discotic Compounds 73 and 74 display electrochemical properties
liquid-crystalline Pcs (71 and 72) wherein TTF subunits are arising from both the TTF subunits and the phthalocyanine
attached to the periphery of the central macrocyclic core by long cores. 1H NMR and UV−vis−NIR spectroscopic studies
and flexible alkyl chains.208 Compounds 71 and 72 were analyzed provided evidence of their strong solvent-dependent aggregation
using polarized light microscopy and differential scanning behavior. For instance, a hypsochromic shift in the Q-bands and a
calorimetry. It was observed that compound 71, with a single broadening of the associated peaks was seen as the temperature
TTF unit, exhibits a mesophase, tentatively assigned as Dhd based was lowered to 0 °C. This is consistent with the formation of
2694 DOI: 10.1021/acs.chemrev.6b00375
Chem. Rev. 2017, 117, 2641−2710
Chemical Reviews Review

Scheme 33. Synthetic Approaches to Symmetrical TTF-Annulated Phthalocyanines

cofacially aggregated Pcs at these lower temperatures. Upon geometries. The authors demonstrated that the emission
addition of iodine to the Pc-TTF constructs, broad CT bands properties of the SiPc core are quenched to varying degrees by
were observed at ≥800 nm, a finding considered to be consistent the axially coordinated electron-rich TTF subunits. The
with the formation of a TTF•+ radical species. Quenching of the magnitude of the quenching depends on both the relative
Pc-based emission was observed at ambient temperature; separation of the two redox-active components (Pc and TTF)
presumably, this reflects CT interactions within these systems and the flexibility of the linking group. In the case of 76b
that serve to deactivate the excited state. (containing a highly flexible linker group) and 76e (with a
Similar kinds of D−σ−A Pcs, bearing 4,5-ethylenedithio-TTF dendritic axial ligand), the emission intensities of the macrocyclic
(EDT-TTF) groups attached to the isoindoline subunits, were cores were reduced by 99% and 96%, respectively, relative to that
reported by Shen and co-workers (75a−c).211−214 The synthetic of a similar SiPc reference compound, W, lacking the TTF units.
procedure is shown in Scheme 36B. The poor solubility of the Molecular modeling studies of these compounds provided a
metal complexes complicated electrochemical analyses. Never- framework for interpreting the degree of emission quenching in
theless, it was found that these Pc species exhibit two distinct terms of the intramolecular spacing. Varying the length of the
redox waves corresponding to oxidation of the TTF units. The axial ligands in these complexes allowed the degree of quenching
TTFs were found act as independent redox centers, as inferred of these TTF-functionalized Pcs to be tuned between 73% and
from the lack of electrochemical involvement of the Pc core <1% relative to that of the reference system W. The strong
under the conditions of experiment. quenching observed in certain case s is consistent with facile and
A series of silicon phthalocyanine (SiPc) bis-esters was rapid photoinduced electron transfer occurring within what are
reported by Beeby and co-workers. These systems, compounds formally coordination-based D−A ensembles.
76a−e, rely on the use of two phenyl carboxylate ligands to ligate The emission behavior of complexes 76 proved highly
the TTF units to the core Pc Si4+ center by axial coordination.215 temperature-dependent, with a sharp emission enhancement
Variations in the substitution patterns around the central being seen at the glass point of the solvent. This latter
aromatic “hinge” within the ligands leads to different molecular observation was ascribed to the axial ligands being frozen into
2695 DOI: 10.1021/acs.chemrev.6b00375
Chem. Rev. 2017, 117, 2641−2710
Chemical Reviews Review

Figure 63. (a) UV−vis−NIR spectra of 61d recorded as a 2 × 10−6 M


solution in THF. (Note: The neutral form is shown with a line, whereas
a dotted line is used for the oxidized form.) (b) Representative thin-layer
CV of 61a in 0.1 M TBAPF6 in CH2Cl2/CH3CN (9.5:0.5 v/v) at a scan
rate of 100 mV s−1 using Pt electrodes. (Reproduced with the permission Figure 64. (a) UV−vis−NIR spectra of compound 64 in CH2Cl2
from ref 195. Copyright 2005 American Chemical Society.) (c) UV− recorded before and after the addition of 1 equiv of TCNQ. (b) Change
vis−NIR spectra of compound 63b recorded in the absence and in the DPV spectra of 64 recorded in CH2Cl2/MeCN (3:2, v/v)
presence of I2. (This figure was redrawn from data that were originally observed upon the incremental addition of sodium perchlorate.
published in ref 199.) (Reproduced with permission from ref 201. Copyright 2008 Elsevier
Ltd.)

conformations wherein PET is blocked. PET could also be


blocked by oxidizing the TTF subunits, leading to the fluorescent
features being turned on. Support for the specific oxidation of the

Table 11. Electrochemical Data for Various TTF-Annulated Phthalocyanines


compd Ered3 (V) Ered2 (V) Ered1 (V) Eox1 (V) Eox2 (V) Eox3 (V) supporting electrolyte solvent ref
61aa
− − −0.98 +0.11 +0.62 − TBAPF6 DCM/MeCN 195
61ba − − − +0.14 +0.60 − TBAPF6 DCM/MeCN 195
61ca − − − +0.14 +0.60 − TBAPF6 DCM/MeCN 195
61da − − − +0.14 +0.59 − TBAPF6 DCM/MeCN 195
62ab − − − +0.16 +0.57 − TBAPF6 DCM 196
63a − − −1.29c +0.43d +0.73 − TBAClO4 DCM 198
63b − −1.36 −1.14 +0.39d +0.76 − TBAClO4 DCM 198
63ce − −1.48 −1.19 +0.42d +0.83 − TBAClO4 DCM 200
64f − − − +0.264 +0.592 − TBAClO4 DCM/MeCN 201
65bg − − − +0.75 +1.25 − TBAPF6 − 202
66ae − −1.39 −1.04 +0.41 +1.36 − TBAClO4 DCM 205
66b,ce − −1.30 −1.03 +0.38 +0.77 − TBAClO4 DCM 205
66de − −1.05 −0.79 +0.38 +0.73 − TBAClO4 DCM 205
67ae − −1.52 −1.18 +0.34d +0.69 − TBAClO4 DCM 207
67be − −1.42 −1.10 +0.44d +0.73 − TBAClO4 DCM 207
67ce − −1.42 −1.10 +0.44d +0.73 − TBAClO4 DCM 207
68ae −1.47d −1.27c −0.78d +0.41d +0.72d − TBAClO4 DCM 199
68be − −1.36c −0.98d +0.39d +0.73d − TBAClO4 DCM 199
69ae − −1.52 −1.18 +0.34d +0.69d − TBAClO4 DCM 199
69be − −1.75 −1.31 +0.33d +0.84d − TBAClO4 DCM 199
70e − − − +0.40 +0.60 +0.81 TBAClO4 DCM 204

a
Electrochemical potentials were measured in CH2Cl2/MeCN (9.5:0.5 v/v) using a Pt working electrode, and a scan rate of 100 mV s−1. Potentials
are in volts vs the Fc/Fc+ couple. bElectrochemical potentials were measured using a Pt electrode as the counter electrode. Potentials are in volts vs
Ag/AgCl. The scan rate was 100 mV s−1. cIrreversible. dQuasireversible. eElectrochemical measurements were carried out using a Ag/AgNO3 (0.01
M) reference electrode, a glassy carbon working electrode, and a Pt counter electrode at a scan rate of 200 mV s−1. fElectrochemical potentials were
recorded in CH2Cl2/MeCN (3:2, v/v) using a 5 × 10−4 M solution of 64. gElectrochemical measurements were carried out using a Ag/AgI reference
electrode and Pt working and counter electrodes. A scan rate of 100 mV s−1 was used.

2696 DOI: 10.1021/acs.chemrev.6b00375


Chem. Rev. 2017, 117, 2641−2710
Chemical Reviews Review

Scheme 34. Synthetic Scheme Leading to Symmetrical TTF-Crown-Ether-Functionalized Phthalocyanines

Chart 8. Chemical Structures of Some Representative


Unsymmetrical TTF-Annulated Phthalocyanines

Figure 65. (a) UV−vis spectral change in the Q-bands of compound


65a observed upon the incremental addition of MeOH into initial
CHCl3 solutions. (b) TEM image of the gelated state of 65a showing its
fibrous structure. (c) TEM image of 65a showing the equal distribution
of both left-hand and right-hand helices. These helices have nanometer
widths and micrometer lengths. Scale bar = 200 nm. (Reproduced with
permission from ref 202. Copyright 2005 Royal Society of Chemistry.)

appended TTF subunits came from spectroelectrochemical


studies. Electrochemical data for all of the compounds discussed
in this section are summarized in Table 12. condensations, the products obtained often exhibit unique
2.8.5. Norphthalocyanines Attached to TTFs through photophysical properties, including a strongly split Q-band
Saturated Spacers. TTF-Crown-Ether-Fused Norphthalocya- ascribed to π−π* transitions.
nines. Tribenzoporphyrazines, or norphthalocyanines, are a class The introduction of a TTF unit onto the periphery of a
of low-symmetry tetrapyrrolic macrocycles216−218 that represent norphthalocyanine core imparts interesting optical and electro-
a special case of AB3 azaporphyrinic macrocyclic compounds in chemical properties that are presumably due to the combined
which A represents a pyrrolic subunit and B is an isoindole presence of a strong chromophore and a redox-active species
moiety. These products are typically obtained by means of within a single molecular skeleton. Hou and co-workers reported
crossover cyclotetramerizations involving the use of phthaloni- an AB3-type of norphthalocyanine (77a) and its Mg2+ complex
trile and maleonitrile precursors. Because of the two different (77b), each of which bears a single TTF unit linked by a thia-
types of peripheral functionalities that result from these mixed crown-ether spacer.219 To our knowledge, 77a and 77b represent
2697 DOI: 10.1021/acs.chemrev.6b00375
Chem. Rev. 2017, 117, 2641−2710
Chemical Reviews Review

Scheme 35. Syntheses of Unsymmetrical TTF-Annulated Phthalocyanines

The ability of 77b to function as a potential donor for


F4TCNQ was demonstrated by the formation of a CT complex,
as evidenced by steady-state absorption, EPR, and IR
spectroscopic studies. Theoretical calculations provided further
support for the proposed intramolecular electronic interactions
within the TTF-bearing molecular construct (Figure 67).
Norphthalocyanines Connected to TTFs through Saturated
Thia-Ether Bridges. Compounds 78a and 78b, reported by Hou
et al., are norphthalocyanines bearing TTF subunits attached by
ethylenedithio spacers.220 The Mg(II) complex, 78b, was
synthesized by carrying out a mixed condensation between an
excess of phthalonitrile and precursor Y in n-butanol at reflux, in
Figure 66. UV−vis−NIR spectral changes observed when the neutral accord with the conventional Mg(II)-template strategy. The free-
form of compound 66a is titrated with TFA in CHCl3. (Reproduced base form, 78a, was prepared by treating 78b with acid followed
with permission from ref 205. Copyright 2011 John Wiley & Sons.)
by an aqueous workup (Scheme 38).
The optical spectra of the unsymmetrical norphthalocyanines
the only examples of TTF-annulated norphthalocyanines (NPcs) 78 recorded in CH2Cl2 are characterized by split Q-
reported to date. Their synthesis is shown in Scheme 37 and bands that obey the Lambert−Beer law up to concentrations of
involves as the key step the Mg2+-templated cross-cyclization of a ca. 10 μM (Figure 68a). This finding was taken as evidence that
thia-crown-ether-based malonitrile precursor, X, with phthaloni- these compounds do not aggregate appreciably in CH2Cl2, THF,
trile at high temperature. or DMF at low concentrations; presumably, this reflects the long
2698 DOI: 10.1021/acs.chemrev.6b00375
Chem. Rev. 2017, 117, 2641−2710
Chemical Reviews Review

Chart 9. Phthalocyanine Derivatives with Peripherally Tethered or Axially Coordinated TTF Units

Scheme 36. Synthetic Routes to Compounds 73a,b and 75b,c

To understand the electron-donor properties of these


alkyl (C12) chains present on the TTF units that serve to block compounds, electrochemical investigations were carried out in
intermolecular interactions. mixed solvents (CH2Cl2/MeCN, 9:1 v/v). The free-base form,
2699 DOI: 10.1021/acs.chemrev.6b00375
Chem. Rev. 2017, 117, 2641−2710
Chemical Reviews Review

Table 12. Electrochemical Data for Phthalocyanines (V vs Fc/Fc+) Bearing Peripherally Fused or Axially Coordinated TTF Units
compd Ered2 (V) Ered1 (V) Eox1 (V) Eox2 (V) Eox3 (V) supporting electrolyte solvent scan rate mV/s ref
71b
− − +0.57 +0.73 +0.99 TEAClO4 DCM 100 208
72b − − +0.56 +0.70ir +0.92 TEAClO4 DCM 100 208
73ab −1.08 −0.74 +0.51 +0.83 − TEAPF6 DMF 1100 209
73bb − −0.51 +0.50 +0.87 +1.18 TBAPF6 BCN 100 210
74b − −0.50 +0.54 +0.92 +1.15 TBAPF6 DCM 100 210
75bc − − +0.620 +0.903 − TEABF4 DCM 12.5 214
75cc − − +0.612 +0.884 − TEABF4 DCM 12.5 214
76ab −1.36 −0.98 +0.23 +0.57 +0.78 TMABF4 DCM 20 215
77bd −1.42 −0.95 +0.47 +0.93 +1.59 TBAPF6 DCM/MeCN 100 219
78ae −1.566 −0.930 +0.476 +0.875 − TBAPF6 DCM/MeCN 100 220
78be −1.435 −1.000 +0.475 +0.863 +1.386 TBAPF6 DCM/MeCN 100 220
a
All measurements were carried out in the presence of 0.1 M supporting electrolyte as indicated and in the noted solvents under an Ar atmosphere.
b
Potentials are V vs the Ag/Ag+ reference and were recorded using Pt working and counter electrodes. cPotentials are reported against SCE.
d
Measurements were carried out in a 7:3 (v/v) mixture of CH2Cl2 and MeCN. ePotentials are reported against SCE in a mixture of CH2Cl2 and
MeCN (9:1 v/v). Note: TMA+ = tetramethylammonium cation.

Scheme 37. Synthesis of Unsymmetrical TTF-Crown-Ether-Fused Norphthalocyanines

properties of SubPcs bear analogy to those of Pcs but show


differences that might be expected given their relatively smaller
14 π-electron conjugation circuits. For instance, SubPcs typically
display an intense absorption band in the visible region (∼550
nm). The inherently bowl-shaped and C3 symmetric structures of
nonfunctionalized SubPcs allows a fine-tuning of their photo-
physical properties by varying their axial ligands. Modifying the
peripheral substituents provides another means of modulating
the basic properties of SubPcs.
Figure 67. (a) HOMO and (b) LUMO contour surfaces of 77b. Recognizing the potential benefits that could come from
(Reproduced with permission from ref 219. Copyright 2015 Elsevier functionalization of the SubPc core, Shimizu et al. reported a
Ltd.) series of TTF-annulated SubPcs.222 These products (79−81)
were synthesized through a conventional boron-templated
reaction involving tetrafluorophthalonitrile and a TTF-fused
78a, exhibits complex electrochemical features dominated by two phthalonitrile bearing thiomethyl substituents (Scheme 39).
reduction signals ascribed to the norphthalocyanine core All three possible TTF-annulated SubPcs (i.e., bearing one,
(NPc−3/NPc−4 and NPc−2/NPc−3) and two oxidation signals two, or three annulated TTF substituents) were found to display
assigned to the TTF units (TTF•+/TTF and TTF2+/TTF•+) features consistent with intramolecular charge transfer in analogy
(Table 12). However, in the case of the Mg(II) complex (78b), a to what was observed in the case of TTF-annulated Pcs as
total of five redox signals are seen: two were assigned (Figure discussed earlier in this review. All three TTF-annulated SubPcs
68b) to NPc-based reductions 78a and three to oxidations, exhibit intense single Q-band absorption with a typical SubPc-
assigned by the authors as TTF•+/TTF, TTF2+/TTF•+ and like shoulder that gradually broadens as the number of fused TTF
NPc−1/NPc−2). subunits increases from the mono-TTF SubPc, 79, through to
2.9. TTF-Annulated Subphthalocyanines the triply fused TTF-SubPc, 81. Moreover, the Q-band gradually
Subphthalocyanine (SubPcs) are ring-contracted phthalocyanine shifts to lower energy, from 578 nm in the case of 79, to 584 nm
analogues that contains a central coordinated boron atom. in the case of 80, and to 593 nm in the case of 81. Fluorescence
SubPcs are synthesized readily and in good yields (in some cases spectral studies of these TTF-SubPcs revealed a monotonic
up to ∼90%) through the cyclotrimerization of phthalonitrile in quenching of the fluorescence intensity as the number of TTF
the presence of a boron trihalide reagent.221 The photophysical units annulated to the SubPc core increased. The quantum yield
2700 DOI: 10.1021/acs.chemrev.6b00375
Chem. Rev. 2017, 117, 2641−2710
Chemical Reviews Review

Scheme 38. Synthesis of 78a,b

∼650−1200-nm spectral range (Figure 69). As invoked in the


case of other TTF-functionalized chromophores, this relative
increase in fluorescence was ascribed to oxidation of the TTF
units and a commensurate decrease in their ability to act as
electron donors.
Electrochemical investigations of these TTF-annulated
SubPcs carried out in o-dichlorobenzene revealed two oxidation
potentials (vs Fc0/Fc+) for each species. The integrated
intensities of the individual redox waves gradually increased as
the number of TTF substituents attached to the SubPc core
increased (Table 13). This finding is consistent with the notion
that these oxidation processes are largely TTF-centered.

3. CONCLUSIONS AND FUTURE PROSPECTS


The parent TTF-annulated pyrroles were first reported 17 years
ago. Studies of TTF-fused oligopyrrolic macrocyclic systems
started shortly thereafter and then blossomed rapidly. Currently,
a large variety of pyrrolo-TTF macrocyles are known, including
TTF-calixpyrroles, TTF-porphyrins, TTF-phthalocyanines,
TTF-subphthalocyanines, and TTF-porphyrazines, to name a
few prominent systems. In this review, we have discussed how the
photophysical and electrochemical sensing behaviors of the core
systems are affected by annulation or tethering of TTF
substituents. Typically, substantial differences in the charge-
transfer (CT) behavior, the electrochemical properties, and the
excited-state dynamics are observed. In some cases, the effects of
TTF functionalization on the metal-free and corresponding
Figure 68. (a) UV−vis−NIR spectral changes observed upon varying metal-complexed forms differ significantly. This subtlety has
the concentration of 78b in CH2Cl2. Inset: Plot of the Q-band allowed for control over both the ground- and excited-state
absorption intensity at 697 nm vs concentration of 78b. (b) CV of 78b. properties and for the development of new ET systems. Anion-
Inset: DPV of 78b recorded in 0.1 M TBAPF6 in a CH2Cl2/MeCN induced conformational changes and cooperative guest-binding
mixture (9:1, v/v). (This figure was redrawn from data that were properties of TTF-pyrrolic macrocycles (i.e., calix[n]pyrrolic
originally published in ref 220.) hybrids) have also been exploited as a means of controlling the
electronic features of TTF-based macrocyclic systems. This
(Φf) determined for the mono-TTF-annulated system, 79, is control, in turn, has allowed for the development of sensors and
0.06. In contrast, the doubly fused SubPc system, 80, has a has provided a means to modulate the nature and extent of
fluorescence quantum yield (Φf) of 6.1 × 10−4, whereas no electron transfer.
detectable fluorescence emission was seen in the case of the tris- Of particular interest are TTF-based systems bearing photo-
TTF-annulated system, 81. The fluorescence quenching was responsive chromophores, such as porphyrins, porphyrazines,
ascribed to the same intramolecular CT mechanism invoked for and phthalocyanines. Several of these constructs have demon-
other TTF-functionalized systems discussed in this review. strated utility in terms of effecting solar energy conversion or
Chemical oxidation using Fe(ClO4)3 led to a concentration- capturing and storing charges at the molecular level. The
dependent increase in the intensity of the CT band located in the progress made has obvious ramifications in terms of mimicking
2701 DOI: 10.1021/acs.chemrev.6b00375
Chem. Rev. 2017, 117, 2641−2710
Chemical Reviews Review

Scheme 39. Synthesis of TTF-Annulated Subphthalocyanines

derivatives could emerge as potential redox mediators in


nonaqueous Li−O2 batteries and might permit enhanced
recharge kinetics and greater cycle reversibility. Here, the ion-
responsive nature of many of the TTF-containing oligopyrrolic
constructs detailed in this review could contribute to effective ion
and charge transport. It is thus our prediction that materials
based on TTF-containing oligopyrrole precursors will continue
to enhance our fundamental understanding while driving
advances across a number of application areas. It is our hope
that the present review, which was deigned to summarize the
current state of the art, will inspire efforts along both of these
lines.

Figure 69. UV−vis−NIR spectral changes seen when the TTF- AUTHOR INFORMATION
annulated SubPc 81 is subjected to chemical oxidation by Fe(ClO4)3 in a
CHCl3/MeCN mixture (2:1 v/v). (Reproduced with permission from Corresponding Authors
ref 222. Copyright 2013 Wiley-VCH Verlag GmbH.) *E-mail: joj@sdu.dk. Tel.: +45-6550-2587.
*E-mail: sessler@cm.utexas.edu. Tel.: +1-512-471-5009.
Table 13. Electrochemical Data for TTF-Annulated
Notes
Subphthalocyanines
The authors declare no competing financial interest.
compd Ered1 (V) Eox1 (V) Eox2 (V)
79 −1.49 +0.20 +0.48 Biographies
80 −1.56 +0.22 +0.43
Atanu Jana received his Ph.D. in Supramolecular Chemistry from the
81 −1.43 +0.16 +0.39
a
Indian Institute of Technology Kanpur, India, in 2010 under the
Measurements were carried out in o-dichlorobenzene containing 0.1 supervision of Professor Parimal K. Bharadwaj. His first postdoctoral
M TBAClO4 as the supporting electrolyte under an Ar atmosphere. research experience was with Professor Jonathan L. Sessler at Yonsei
Potentials are in volts vs the Fc/Fc+ couple recorded under the
University (Seoul, South Korea) under the support of a World Class
conditions of analysis. The scan rate was 100 mV s−1.
University (WCU) research program led by Prof. Dongho Kim. After his
time in Seoul, Dr. Jana moved to the National Institute for Materials
artificial photosynthesis and advancing solar cell applications. Science (NIMS), Tsukuba, Japan, and worked with Dr. Jonathan P. Hill
TTF-functionalized oligopyrrole systems have also proved useful and Professor Katsuhiko Ariga before joining Professor Michael D.
in terms of supporting self-assembly and effecting sensing,
Ward’s Group at the University of Sheffield, U.K., as a European
particularly for electron-deficient substrates. In both of these
Commission Marie Curie International Incoming Research Fellow
broad application areas, challenges remain. For instance,
(2013−2015). His current research interests focus on various
improvements in solar cell performance could be envisioned if
macrocyclic materials constructed using, among others, crown ether,
one could learn to control more precisely the dimensional
cryptand, TTF-calixpyrrole, and TTF-porphyrin building blocks, as well
engineering of well-defined self-assembled nanoarchitectures by
as studies of their properties, supramolecular charge-transfer phenom-
using TTF-based interactions to align molecular dipoles, to
ena, guest recognition, and biological and medicinal applications.
control more precisely molecular electronics, and to establish
and maintain orbital symmetries. Separately, an ability to Masatoshi Ishida received his Ph.D. degree from Kyushu University in
engineer the controlled disassembly of aggregated TTF- 2010 under the supervision of Professor Yoshinori Naruta. After
pyrrole-based systems could allow for the construction of ever working as a research fellow at the Institute for Materials Chemistry and
more sensitive sensors for nitroaromatic explosives, as well as Engineering (IMCE), Kyushu University, he worked with Professors
potentially a range of other substrates. Systems containing Jonathan L. Sessler and Dongho Kim as a WCU postdoctoral research
multiple TTF units could also emerge as promising charge- fellow at Yonsei University (Seoul, South Korea). He is currently an
storage devices and might lead to new types of electrode Assistant Professor in the Department of Chemistry and Biochemistry
materials suitable for use in batteries. In particular, TTF and Center for Molecular Systems (CMS) at Kyushu University

2702 DOI: 10.1021/acs.chemrev.6b00375


Chem. Rev. 2017, 117, 2641−2710
Chemical Reviews Review

(Fukuoka, Japan). His current research interests are focused on the Republic of Korea. S.B. gratefully acknowledges The Danish
synthetic and redox chemistry of novel π-conjugated porphyrinoids. Council for Independent Research, Technology, and Production
Jung Su Park received his Ph.D. from The University of Texas at Austin Sciences (FTP, Project 5054-00052). J.O.J. thanks the Villum
in 2010 under the supervision of Professor Jonathan L. Sessler. He then Foundation and the Danish Natural Science Research Council
worked with Professor Chad Mirkin as a postdoctoral research fellow at (FNU, Project 11-106744) for financial support. J.L.S. acknowl-
Northwestern University (Evanston, IL). He is presently an Assistant edges the U.S. National Science Foundation (Grant CHE-
Professor in the Department of Chemistry at Sookmyung Women’s 1402004), the Robert A. Welch Foundation, and the Korean
University (Seoul, South Korea). His current research focus involves the World Class University Program for support of this work.
development of responsive organic functional materials based on redox-,
photo-, and chemically responsive organic building blocks. ABBREVIATIONS
Steffen Bähring received his Ph.D. in Supramolecular Chemistry from
BArF− tetrakis[bis(3,5-trifluoromethyl)phenyl]borate)
anion
the University of Southern Denmark (Odense, Denmark) in 2013 under
BHEEN 1,5-bis[(2-hydroxyethoxy)ethoxy]naphthalene
the guidance of Professors Kent A. Nielsen and Jan O. Jeppesen. While
BIQ bisimidazolium quinine (1,3,5,7-tetramesityl-4,8-
working on molecular sensors incorporating tetrathiafulvalene-calix[4]-
dioxo-3,4,7,8-tetrahydrobenzo[1,2-d:4,5-d′]-
pyrrole, he joined the group of Professor Jonathan L. Sessler for a
diimidazole-1,5-diium)
research stay during 2011−2012. After the awarding of his Ph.D., he
BTTF-C4P benzo-TTF-calix[4]pyrrole
worked at the Danish Technological Institute. During this period, he
C[4]P calix[4]pyrrole
focused on supporting research and development efforts for Danish
CD β-cyclodextrin
industry, with a particular emphasis on oil and gas. He is currently a
CHCl3 chloroform
Danish Independent Postdoctoral Researcher in the groups of
CPK Cory−Pauling−Koltun
Professors Jeppesen, Sessler, and Guldi, where he is working on organic
CS charge-separated
photovoltaics based on tetrathiafulvalene.
CT charge-transfer
Jan O. Jeppesen received his B.Sc., M.Sc., and Ph.D. degrees in 1996, CV cyclic voltammogram
1999, and 2001, respectively, from the University of Southern Denmark C60 fullerene
(Odense, Denmark) for work in supramolecular tetrathiafulvalene CBPQT4+ cyclobis(paraquat-p-phenylene)
chemistry. In the first part of his Ph.D. studies, he developed a synthetic DBU 1,8-diazabicyclo[5.4.0]undec-7-ene
route to pyrrolo-tetrathiafulvalenes, and in 2000−2001, he joined the DCE 1,2-dichloroethane
group of Professor J. Fraser Stoddart at the University of California, Los DCM dicholoromethane
Angeles (UCLA), working on the design and synthesis of amphiphilic DDQ 2,3-dichloro-5,6-dicyano-1,4-quinolene
bistable [2]rotaxanes. He is the recipient of several research prizes DFT density functional theory
including the Ellen and Niels Bjerrum Chemistry Award (2014), the DLS dynamic light scattering
Torkil Holm prize (2010), the Villum Kann Rasmussens award (2009), DMSO dimethyl sulfoxide
the Aksel Tovborg Jensens award (2009), the Bjerrum−Brøndsted− DMF N,N-dimethylformamide
Lang lectureship (2009), and the Hede Nielsens prize (2003). He is a DNP-C4P bis-dinitrophenyl-calix[4]pyrrole
member of the Executive Committee of The Danish Council for DNT 2,4-dinitrotoluene
Independent Research, Technology, and Production Sciences (FTP). DONP dioxynaphthalene
Since 2008, Jan O. Jeppesen has been a Full Professor at the University DOSY-NMR diffusion-ordered NMR
of Southern Denmark, and he is one of the world’s leading scientists in DPV differential pulse voltammetry
the areas of synthetic and supramolecular chemistry of tetrathiafulva- DVD digital versatile disc
lenes. Home page: http://www.jojgroup.sdu.dk. EDT-TTF 4,5-ethylenedithiotetrathiafulvalene
Jonathan L. Sessler received a B.Sc. degree in Chemistry in 1977 from EPR electron paramagnetic resonance
the University of California, Berkeley. He obtained his Ph.D. from ET electron transfer
Stanford University in 1982. He was an NSF-NATO and NSF-CNRS F4TCNQ 2,3,5,6-tetrafluoro-7,7,8,8-tetracyanoquinodime-
Postdoctoral Fellow at Université Louis Pasteur de Strasbourg (1982− thane
1983). In 1984, he accepted a position as an Assistant Professor of HOMO highest occupied molecular orbital
Chemistry at the University of Texas at Austin, where he is currently the ICT intramolecular charge transfer
Doherty-Welch Chair. Professor Sessler was also a WCU Professor at IPCE incident photon-to-current efficiency
Yonsei University in Seoul, South Korea. He recently accepted a IR infrared
summer research professorship and laboratory directorate at Shanghai ITC isothermal titration calorimetry
University. Professor Sessler is one of the world’s leading scientists LB Langmuir−Blodgett
working in the areas of supramolecular chemistry and expanded LUMO lowest unoccupied molecular orbital
porphyrin chemistry. Home page: http://sessler.cm.utexas.edu/Home. MALDI-TOF matrix-assisted laser desorption ionization time-
html. of-flight
MeCN acetonitrile
MO molecular orbital
ACKNOWLEDGMENTS MSA methanesulfonic acid
A.J. gratefully acknowledges a Marie Curie International NDI naphthalenediimide carboxylate
Incoming Fellowship (MC-IIF). M.I. acknowledges a JSPS NIR near-infrared
Grant-in-Aid for scientific research (No. 16K05700). J.S.P. is NMR nuclear magnetic resonance
grateful for support under a “Cooperative Research Program for NMF N-methylfulleropyrrolidine
Agriculture Science & Technology Development (Project NPc norphthalocyanine
PJ010506)” grant from the Rural Development Administration, OTE optically transparent electrode
2703 DOI: 10.1021/acs.chemrev.6b00375
Chem. Rev. 2017, 117, 2641−2710
Chemical Reviews Review

PCBA phenyl-C61-butyric acid (6) Peterson, I. R. In Nanostructures Based on Molecular Materials;


Pc phthalocyanine Göpel, W., Ziegler, C.; Eds.; VCH Publishers: Weinheim, Germany,
PET photoinduced electron transfer 1992; p195.
PhCN benzonitrile (7) Ashwell, G. J.; Sage, I.; Trundle, C. In Molecular Electronics;
PNC porphyrinic nanochannel Ashwell, G. J., Ed.; Wiley: New York, 1992; Chapter 1.
(8) Metzger, R. M. The Quest for D−σ−A Unimolecular Rectifiers and
ppb parts per billion
Related Topics in Molecular Electronics. In Molecular and Biomolecular
Pz porphyrazine Electronics; Birge, R., Ed.; Advances in Chemistry Series; American
SAM self-assembled monolayer Chemical Society: Washington, DC, 1994; Vol. 240, Chapter 5, pp 81−
SCE secondary calomel electrode 129.
SEM scanning electron microscopy (9) Metzger, R. M. D−σ−A Unimolecular Rectifiers. Mater. Sci. Eng., C
SiPc silicon phthalocyanine 1995, 3, 277−285.
SOMO singly occupied molecular orbital (10) Bloor, D. In Introduction to Molecular Electronics; Petty, M. C.,
SubPc subphthalocyanine Bryce, M. R., Bloor, D., Eds.; Oxford University Press: Oxford, U.K.,
SWV square-wave voltammetry 1995; Chapter 1.
TBAB tetrakis-n-butylammonium bromide (11) Metzger, R. M.; Chen, B.; Höpfner, U.; Lakshmikantham, M. V.;
TBACl tetrabutylammonium chloride Vuillaume, D.; Kawai, T.; Wu, X.; Tachibana, H.; Hughes, T. V.; Sakurai,
TBAClO4 tetrabutylammonium perchlorate H.; et al. Unimolecular Electrical Rectification in Hexadecylquinolinium
TBAH tetrakis-n-butylammonium hydroxide Tricyanoquinodimethanide. J. Am. Chem. Soc. 1997, 119, 10455−10466.
TBAHP tetrakis-n-butylammonium hexafluorophosphate (12) Khodorkovsky, V.; Becker, J. Y. Molecular Design of Organic
Conductors. In Organic Conductors: Fundamentals and Applications;
TBAI tetrabutylammonium iodide
Farges, J.-P., Ed.; Marcel Dekker: New York, 1994; Chapter 3, pp 75−
TBAP tetrakis-n-butylammonium perchlorate 88.
TBAPF6 tetrakis-n-butylammonium hexafluorophosphate (13) Imahori, H.; Sakata, Y. Donor-Linked Fullerenes: Photoinduced
TCNQ 7,7,8,8-tetracyano-p-quinodimethane Electron Transfer and Its Potential Application. Adv. Mater. 1997, 9,
TDDFT time-dependent density functional theory 537−546.
TEABArF tetraethylammonium tetrakis[bis(3,5- (14) Li, F.; Gentemann, S.; Kalsbeck, W. A.; Seth, J.; Lindsey, J. S.;
trifluoromethyl)phenyl]borate Holten, D.; Bocian, D. F. Effects of Central Metal Ion (Mg, Zn) and
TEACl tetrakis-ethylammonium chloride Solvent on Singlet Excited-State Energy Flow in Porphyrin-Based
TEAI tetraethylammonium iodide Nanostructures. J. Mater. Chem. 1997, 7, 1245−1262.
TEM transmission electron microscopy (15) Vollmer, M. S.; Würthner, F.; Effenberger, F.; Emele, P.; Meyer,
TFA trifluoroacetic acid D. U.; Stümpfig, T.; Port, H.; Wolf, H. C. Anthryloligothienylporphyr-
THACl tetrakis-hydroxyammonium chloride ins: Energy Transfer and Light-Harvesting Systems. Chem. - Eur. J. 1998,
THF tetrahydrofuran 4, 260−269.
(16) Marder, S. R.; Kippelen, B.; Jen, A. K. Y.; Peyghambarian, N.
TMPyP 5,10,15,20-tetrakis(N-methyl-4-pyridinio)-
Design and Synthesis of Chromophores and Polymers For Electro-
porphyrin
Optic and Photorefractive Applications. Nature 1997, 388, 845−851.
TNB 1,3,5-trinitrobenzene (17) Verbiest, T.; Houbrechts, S.; Kauranen, M.; Clays, K.; Persoons,
TNDCF glycol diester-linked bis-2,5,7-trinitrodicyanome- A. Second-Order Nonlinear Optical Materials: Recent Advances in
thylenefluorene-4-carboxylate Chromophore Design. J. Mater. Chem. 1997, 7, 2175−2189.
TNP 2,4,6-trinitrophenol (picric acid) (18) Gordon, P. F.; Gregory, P. Organic Chemistry in Colour; Springer-
TNT 2,4,6-trinitrotoluene Verlag: Berlin, 1987.
TPA two-photon absorption (19) Goetz, K. P.; Vermeulen, D.; Payne, M. E.; Kloc, C.; McNeil, L. E.;
TPP 5,10,15,20-tetraphenylporphyrin Jurchescu, O. D. Charge-Transfer Complexes: New Perspectives On an
TREPR time-resolved electron paramagnetic resonance Old Class of Compounds. J. Mater. Chem. C 2014, 2, 3065−3076.
TTF tetrathiafulvalene (20) Mulliken, R. S. Molecular Compounds and their Spectra. II. J. Am.
TTFCD TTF-bridged bis(β-cyclodextrin)-porphyrin de- Chem. Soc. 1952, 74, 811−824.
rivative (21) Marcus, R. A.; Sutin, N. Electron Transfers in Chemistry and
Biology. Biochim. Biophys. Acta, Rev. Bioenerg. 1985, 811, 265−322.
Zn(hfac)2 hexafluoroacetylacetonato-zinc(II)
(22) Marcus, R. A. Electron Transfer Reactions in Chemistry: Theory
and Experiment (Nobel Lecture). Angew. Chem., Int. Ed. Engl. 1993, 32,
1111−1121.
REFERENCES (23) Heckmann, A.; Lambert, C. Organic Mixed-Valence Compounds:
(1) Jeppesen, J. O.; Becher, J. Pyrrolo-Tetrathiafulvalenes and Their A Playground for Electrons and Holes. Angew. Chem., Int. Ed. 2012, 51,
Applications in Molecular and Supramolecular Chemistry. Eur. J. Org. 326−392.
Chem. 2003, 2003 (17), 3245−3266. (24) Ohkubo, K.; Fukuzumi, S. Long-Lived Charge-Separated States of
(2) Aviram, A.; Ratner, M. A. Molecular Rectifiers. Chem. Phys. Lett. Simple Electron Donor−Acceptor Dyads Using Porphyrins and
1974, 29, 277−283. Phthalocyanines. J. Porphyrins Phthalocyanines 2008, 12, 993−1004.
(3) Metzger, R. M.; Panetta, C. A. Toward Organic Rectifiers. In (25) Fukuzumi, S.; Ohkubo, K.; Suenobu, T. Long-Lived Charge
Molecular Electronic Devices Carter, F. L., Ed.; Marcel Dekker: New York, Separation and Applications in Artificial Photosynthesis. Acc. Chem. Res.
1987; Vol. II, pp 5−25. 2014, 47, 1455−1464.
(4) Metzger, R. M.; Schumaker, R. R.; Cava, M. P.; Laidlaw, R. K.; (26) Kang, Y. K.; Iovine, P. M.; Therien, M. J. Electron Transfer
Panetta, C. A.; Torres, E. Langmuir-Blodgett Films of Donor Urethane Reactions of Rigid, Cofacially Compressed, π-Stacked Porphyrin−
TCNQ and Related Molecules. Langmuir 1988, 4, 298−304. Bridge−Quinone Systems. Coord. Chem. Rev. 2011, 255, 804−824.
(5) Carter, F. L.; Siatkowski, R. E. In From Atoms to Polymers: (27) Bill, N. L.; Trukhina, O.; Sessler, J. L.; Torres, T. Supramolecular
Isoelectronic Analogies; Liebman, J. F., Greenberg, A., Eds.; Molecular Electron Transfer-Based Switching Involving Pyrrolic Macrocycles. A
Structure & Energetics Series; Wiley-VCH: Weinheim, Germany, 1989; New Approach to Sensor Development? Chem. Commun. 2015, 51,
Vol. 11, Chapter 9. 7781−7794.

2704 DOI: 10.1021/acs.chemrev.6b00375


Chem. Rev. 2017, 117, 2641−2710
Chemical Reviews Review

(28) Liu, Y.; Flood, A. H.; Moskowitz, R. M.; Stoddart, J. F. Versatile Switch with Multiple Channel Outputs. J. Am. Chem. Soc. 2011, 133,
Self-Complexing Compounds Based on Covalently Linked Donor− 13256−13259.
Acceptor Cyclophanes. Chem. - Eur. J. 2005, 11, 369−385. (47) Becher, J.; Lau, J.; Mùrk, P. in Electronic Materials: The Oligomer
(29) Zhang, Q.; Tu, Y.; Tian, H.; Zhao, Y.-L.; Stoddart, J. F.; Ågren, H. Approach; Müllen, K., Wegner, G., Eds.; Wiley-VCH: Weinheim,
Working Mechanism for a Redox Switchable Molecular Machine Based Germany, 1998; p 198.
on Cyclodextrin: A Free Energy Profile Approach. J. Phys. Chem. B 2010, (48) Otsubo, T.; Takimiya, K. Recent Synthetic Advances of
114, 6561−6566. Tetrathiafulvalene-Based Organic Conductors. Bull. Chem. Soc. Jpn.
(30) Barin, G.; Coskun, A.; Friedman, D. C.; Olson, M. A.; Colvin, M. 2004, 77, 43−58.
T.; Carmielli, R.; Dey, S. K.; Bozdemir, O. A.; Wasielewski, M. R.; (49) Canevet, D.; Sallé, M.; Zhang, G. X.; Zhang, D. Q.; Zhu, D. B.
Stoddart, J. F. A Multistate Switchable 3 Rotacatenane. Chem. - Eur. J. Tetrathiafulvalene (TTF) Derivatives: Key Building-Blocks for Switch-
2011, 17, 213−222. able Processes. Chem. Commun. 2009, 2245−2269.
(31) Hansen, S. W.; Stein, P. C.; Sørensen, A.; Share, A. I.; Witlicki, E. (50) Fukuzumi, S.; Ohkubo, K.; D’Souza, F.; Sessler, J. L.
H.; Kongsted, J.; Flood, A. H.; Jeppesen, J. O. Quantification of the π−π Supramolecular Electron Transfer by Anion Binding. Chem. Commun.
Interactions that Govern Tertiary Structure in Donor−Acceptor [2] 2012, 48, 9801−9815.
Pseudorotaxanes. J. Am. Chem. Soc. 2012, 134, 3857−3863. (51) Narita, M.; Pittman, C. U., Jr. Preparation of Tetrathiafulvalenes
(32) Coskun, A.; Spruell, J. M.; Barin, G.; Dichtel, W. R.; Flood, A. H.;
(TTF) and Their Selenium Analogs-Tetraselenafulvalenes (TSEF).
Botros, Y. Y.; Stoddart, J. F. High Hopes: Can Molecular Electronics
Synthesis 1976, 1976, 489−514.
Realise Its Potential? Chem. Soc. Rev. 2012, 41, 4827−4859.
(52) Krief, A. Syntheses of Tetraheterofulvalenes and of Vinylene
(33) Spruell, J. M.; Coskun, A.; Friedman, D. C.; Forgan, R. S.;
Sarjeant, A. A.; Trabolsi, A.; Fahrenbach, A. C.; Barin, G.; Paxton, W. F.; Triheterocarbonates − Strategy and Practice. Tetrahedron 1986, 42,
Dey, S. K.; et al. Highly Stable Tetrathiafulvalene Radical Dimers in 1209−1252.
[3]Catenanes. Nat. Chem. 2010, 2, 870−879. (53) Schukat, G.; Richter, A. M.; Fanghänel, E. Synthesis, Reactions,
(34) Coskun, A.; Spruell, J. M.; Barin, G.; Fahrenbach, A. C.; Forgan, R. and Selected Physico-Chemical Properties of 1,3-and 1,2-Tetrachalco-
S.; Colvin, M. T.; Carmieli, R.; Benítez, D.; Tkatchouk, E.; Friedman, D. genafulvalenes. Sulfur Rep. 1987, 7, 155−231.
C.; et al. Mechanically Stabilized Tetrathiafulvalene Radical Dimers. J. (54) Bryce, M. R. Recent Progress on Conducting Organic Charge-
Am. Chem. Soc. 2011, 133, 4538−4547. Transfer Salts. Chem. Soc. Rev. 1991, 20, 355−390.
(35) Azov, V. A.; Gómez, R.; Stelten, J. Synthesis of Electrochemically (55) Schukat, G.; Fanghänel, E. Synthesis, Reactions, and Selected
Responsive TTF-Based Molecular Tweezers: Evidence of Tight Physico-Chemical Properties of 1,3- and 1,2-Tetrachalcogenafulvalenes.
Intramolecular TTF Pairing in Solution. Tetrahedron 2008, 64, 1909− Sulfur Rep. 1993, 14, 245−383.
1917. (56) Jørgensen, T.; Hansen, T. K.; Becher, J. Tetrathiafulvalenes as
(36) Skibiński, M.; Gómez, R.; Lork, E.; Azov, V. A. Redox Responsive Building-Blocks in Supramolecular Chemistry. Chem. Soc. Rev. 1994, 23,
Molecular Tweezers with Tetrathiafulvalene Units: Synthesis, Electro- 41−51.
chemistry, and Binding Properties. Tetrahedron 2009, 65, 10348− (57) Adam, M.; Müllen, K. Oligomeric Tetrathiafulvalenes − Extended
10354. Donors for Increasing the Dimensionality of Electrical-Conduction.
(37) Düker, M. H.; Schäfer, H.; Zeller, M.; Azov, V. A. Rationally Adv. Mater. 1994, 6, 439−459.
Designed Calix[4]arene−Pyrrolotetrathiafulvalene Receptors for Elec- (58) Bryce, M. R. Current Trends in Tetrathiafulvalene Chemistry −
tron-Deficient Neutral Guests. J. Org. Chem. 2013, 78, 4905−4912. Towards Increased Dimensionality. J. Mater. Chem. 1995, 5, 1481−
(38) Jeppesen, J. O.; Nielsen, K. A.; Perkins, J.; Vignon, S. A.; Di Fabio, 1496.
A.; Ballardini, R.; Gandolfi, M. T.; Venturi, M.; Balzani, V.; Becher, J.; (59) Nielsen, M. B.; Becher, J. Two- and Three-Dimensional
et al. Amphiphilic Bistable Rotaxanes. Chem. - Eur. J. 2003, 9, 2982− Tetrathiafulvalene Macrocycles. Liebigs Ann./Recl. 1997, 1997, 2177−
3007. 2187.
(39) Jeppesen, J. O.; Collier, C. P.; Heath, J. R.; Luo, Y.; Nielsen, K. A.; (60) Roncali, J. Linearly Extended π-Donors: When Tetrathiafulvalene
Perkins, J.; Fraser Stoddart, J.; Wong, E. Artificial Molecular Devices Meets Conjugated Oligomers and Polymers. J. Mater. Chem. 1997, 7,
Based on Tetrathiafulvalene. J. Phys. IV 2004, 114, 511−513. 2307−2321.
(40) Jang, Y. H.; Hwang, S. G.; Kim, Y. H.; Jang, S. S.; Goddard, W. A. (61) Bryce, M. R. Tetrathiafulvalenes as π-Electron Donors for
Density Functional Theory Studies of the [2] Rotaxane Component of Intramolecular Charge-Transfer materials. Adv. Mater. 1999, 11, 11−23.
the Stoddart-Heath Molecular Switch. J. Am. Chem. Soc. 2004, 126, (62) Schukat, G.; Fanghänel, E. Synthesis, Reactions, and Selected
12636−12645. Physico-Chemical Properties of 1,3- and 1,2-Tetrachalcogenafulvalenes.
(41) Liu, Y.; Saha, S.; Vignon, S. A.; Flood, A. H.; Stoddart, J. F.
Sulfur Rep. 1996, 18, 1−278.
Template-Directed Syntheses of Configurable and Reconfigurable (63) Garín, J. The Reactivity of Tetrathiaselenafulvalenes and
Molecular Switches. Synthesis 2005, 3437−3445.
Tetraselenafulvalenes. Adv. Heterocycl. Chem. 1995, 62, 249−304.
(42) Chiang, P. T.; Cheng, P. N.; Lin, C. F.; Liu, Y. H.; Lai, C. C.; Peng,
(64) Simonsen, K. B.; Svenstrup, N.; Lau, J.; Simonsen, O.; Mork, P.;
S. M.; Chiu, S. H. A Macrocycle/Molecular-Clip Complex That
Kristensen, G. J.; Becher, J. Sequential Functionalisation of Bis-
Functions as a Quadruply Controllable Molecular Switch. Chem. - Eur. J.
2006, 12, 865−876. Protected Tetrathiafulvalene-Dithiolates. Synthesis 1996, 1996, 407−
(43) Liu, Y.; Wang, C.; Li, M.; Lv, S.; Lai, G.; Shen, Y. A New 418.
Fluorescence Molecular Switch Incorporating TTF and Tetraphenyl- (65) Otsubo, T.; Aso, Y.; Takimiya, K. Dimeric Tetrathiafulvalenes:
porphyrin Units. J. Porphyrins Phthalocyanines 2007, 11, 729−735. New Electron Donors. Adv. Mater. 1996, 8, 203−211.
(44) Tan, W.; Zhang, D.; Wu, H.; Zhu, D. A New 4-(N,N- (66) Simonsen, K. B.; Becher, J. Tetrathiafulvalene Thiolates:
Dimethylamino)benzonitrile (DMABN) Derivative with Tetrathiaful- Important Synthetic Building Blocks for Macrocyclic and Supra-
valene Unit: Modulation of the Dual Fluorescence of DMABN by Redox molecular Chemistry. Synlett 1997, 1997, 1211−1220.
Reaction of Tetrathiafulvalene Unit. Tetrahedron Lett. 2008, 49, 1361− (67) Day, P.; Kurmoo, M. Molecular Magnetic Semiconductors,
1364. Metals and Superconductors: BEDT-TTF salts with magnetic anions. J.
(45) Liu, L.; Zhang, G.; Tan, W.; Zhang, D.; Zhu, D. A Temperature- Mater. Chem. 1997, 7, 1291−1295.
Regulated Molecular Redox Fluorescence Switch Based on a Triad (68) Coronado, E.; Gomez-García, C. J. Polyoxometalate-Based
Bearing Tetrathiafulvalene, Maleimide and Pyrene Moieties. Chem. Phys. Molecular Materials. Chem. Rev. 1998, 98, 273−296.
Lett. 2008, 465, 230−233. (69) Bryce, M. R.; Devonport, W.; Goldenberg, L. M.; Wang, C. S.
(46) Simão, C.; Mas-Torrent, M.; Casado-Montenegro, J.; Otón, F.; Macromolecular Tetrathiafulvalene Chemistry. Chem. Commun. 1998,
Veciana, J.; Rovira, C. A Three-State Surface-Confined Molecular 945−951.

2705 DOI: 10.1021/acs.chemrev.6b00375


Chem. Rev. 2017, 117, 2641−2710
Chemical Reviews Review

(70) Bryce, M. R. Functionalised Tetrathiafulvalenes: New Applica- (93) Hansen, J. A.; Becher, J.; Jeppesen, J. O.; Levillain, E.; Nielsen, M.
tions as Versatile π-Electron Systems in Materials Chemistry. J. Mater. B.; Petersen, B. M.; Petersen, J. C.; Şahin, Y. Synthesis and Non-Linear
Chem. 2000, 10, 589−598. Optical Properties of Mono-pyrrolotetrathiafulvalene Derived
(71) Nielsen, M. B.; Lomholt, C.; Becher, J. Tetrathiafulvalenes as Donor−π−Acceptor Dyads. J. Mater. Chem. 2004, 14, 179−184.
Building Blocks in Supramolecular Chemistry II. Chem. Soc. Rev. 2000, (94) O’Driscoll, L. J.; Andersen, S. S.; Solano, M. V.; Bendixen, D.;
29, 153−164. Jensen, M.; Duedal, T.; Lycoops, J.; van der Pol, C.; Sørensen, R. E.;
(72) Segura, J. L.; Martin, N. New Concepts in Tetrathiafulvalene Larsen, K. R.; et al. Advances in the Synthesis of Functionalised
Chemistry. Angew. Chem., Int. Ed. 2001, 40, 1372−1409. Pyrrolotetrathiafulvalenes. Beilstein J. Org. Chem. 2015, 11, 1112−1122.
(73) Iyoda, M.; Hasegawa, M.; Miyake, Y. Bi-TTF, bis-TTF, and (95) Park, J. S.; Le Derf, F.; Bejger, C. M.; Lynch, V. M.; Sessler, J. L.;
Related TTF Oligomers. Chem. Rev. 2004, 104, 5085−5113. Nielsen, K. A.; Johnsen, C.; Jeppesen, J. O. Positive Homotropic
(74) Yamada, J.; Akutsu, H.; Nishikawa, H.; Kikuchi, K. New Trends in Allosteric Receptors for Neutral Guests: Annulated Tetrathiafulvalene−
the Synthesis of pi-Electron Donors for Molecular Conductors and Calix[4]pyrroles as Colorimetric Chemosensors for Nitroaromatic
Superconductors. Chem. Rev. 2004, 104, 5057−5083. Explosives. Chem. - Eur. J. 2010, 16, 848−854 and references therein..
(75) Fabre, J. M. Synthesis Strategies and Chemistry of Nonsym- (96) Nielsen, K. A.; Jeppesen, J. O.; Levillain, E.; Becher, J. Mono-
metrically Substituted Tetrachalcogenafulvalenes. Chem. Rev. 2004, 104, Tetrathiafulvalene Calix[4]pyrrole in the Electrochemical Sensing of
5133−5150. Anions. Angew. Chem., Int. Ed. 2003, 42, 187−191.
(76) Kobayashi, A.; Fujiwara, E.; Kobayashi, H. Single-Component (97) Nielsen, K. A.; Cho, W.-S.; Lyskawa, J.; Levillain, E.; Lynch, V. M.;
Molecular Metals with Extended-TTF Dithiolate Ligands. Chem. Rev. Sessler, J. L.; Jeppesen, J. O. Tetrathiafulvalene-Calix[4]pyrroles:
2004, 104, 5243−5264. Synthesis, Anion Binding, and Electrochemical Properties. J. Am.
(77) Frére, P.; Skabara, P. J. Salts of Extended Tetrathiafulvalene Chem. Soc. 2006, 128, 2444−2451.
Analogues: Relationships Between Molecular Structure, Electro- (98) Nielsen, K. A.; Cho, W.-S.; Jeppesen, J. O.; Lynch, V. M.; Becher,
chemical Properties and Solid State Organisation. Chem. Soc. Rev. J.; Sessler, J. L. Tetra-TTF Calix[4]pyrrole: A Rationally Designed
2005, 34, 69−98. Receptor for Electron-Deficient Neutral Guests. J. Am. Chem. Soc. 2004,
(78) Lorcy, D.; Bellec, N.; Fourmigue, M.; Avarvari, N. Tetrathia- 126, 16296−16297.
fulvalene-Based Group XV Ligands: Synthesis, Coordination Chemistry (99) Zhu, W.; Park, J. S.; Sessler, J. L.; Gaitas, A. A Colorimetric
and Radical Cation Salts. Coord. Chem. Rev. 2009, 253, 1398−1438. Receptor Combined with a Microcantilever Sensor for Explosive Vapor
(79) Rabaça, S.; Almeida, M. Dithiolene Complexes Containing N Detection. Appl. Phys. Lett. 2011, 98, 123501.
Coordinating Groups and Corresponding Tetrathiafulvalene Donors. (100) Nielsen, K. A.; Stein, P. C. Self-Assembly of Dimeric
Coord. Chem. Rev. 2010, 254, 1493−1508. Tetrathiafulvalene-Calix[4]pyrrole: Receptor for 1,3,5-Trinitrobenzene.
(80) Jeppesen, J. O.; Nielsen, M. B.; Becher, J. Tetrathiafulvalene Org. Lett. 2011, 13, 6176−6179.
Cyclophanes and Cage Molecules. Chem. Rev. 2004, 104, 5115−5131. (101) Nielsen, K. A. A Colorimetric Tetrathiafulvalene-Calix[4]-
(81) Rovira, C. Bis(ethylenethio)tetrathiafulvalene (BET-TTF) and Pyrrole Anion Sensor. Tetrahedron Lett. 2012, 53, 5616−5618.
(102) Nielsen, K. A.; Bähring, S.; Jeppesen, J. O. Acid/Base
Related Dissymmetrical Electron Donors: From the Molecule to
Controllable Molecular Recognition. Chem. - Eur. J. 2011, 17, 11001−
Functional Molecular Materials and Devices (OFETs). Chem. Rev.
11007.
2004, 104, 5289−5317.
(103) Bähring, S.; Olsen, G.; Stein, P. C.; Kongsted, J.; Nielsen, K. A.
(82) Hansen, J. G.; Bang, K. S.; Thorup, N.; Becher, J. Donor−
Coordination-Driven Switching of a Preorganized and Cooperative
Acceptor Macrocycles Incorporating Tetrathiafulvalene and Pyromel-
Calix[4]pyrrole Receptor. Chem. - Eur. J. 2013, 19, 2768−2775.
litic Diimide: Syntheses and Crystal Structures. Eur. J. Org. Chem. 2000, (104) Bosco, F. G.; Bache, M.; Hwu, E.-T.; Chen, C. H.; Andersen, S.
2000, 2135−2144. S.; Nielsen, K. A.; Keller, S. S.; Jeppesen, J. O.; Hwang, I.-S.; Boisen, A.
(83) Simonsen, K. B.; Zong, K. W.; Rogers, R. D.; Cava, M. P.; Becher, Statistical Analysis of DNT Detection Using Chemically Functionalized
J. Stable Macrocyclic and Tethered Donor−Acceptor Systems. Microcantilever Arrays. Sens. Actuators, B 2012, 171−172, 1054−1059.
Intramolecular Bipyridinium and Tetrathiafulvalene Assemblies. J. (105) Nielsen, K. A.; Cho, W.-S.; Sarova, G. H.; Petersen, B. M.; Bond,
Org. Chem. 1997, 62, 679−686. A. D.; Becher, J.; Jensen, F.; Guldi, D. M.; Sessler, J. L.; Jeppesen, J. O.
(84) Bergkamp, J. J.; Decurtins, S.; Liu, S.-X. Current Advances in Supramolecular Receptor Design: Anion-Triggered Binding of C60.
Fused Tetrathiafulvalene Donor−Acceptor Systems. Chem. Soc. Rev. Angew. Chem., Int. Ed. 2006, 45, 6848−6853.
2015, 44, 863−874. (106) Davis, C. M.; Lim, J. M.; Larsen, K. R.; Kim, D. S.; Sung, Y. M.;
(85) Pop, F.; Avarvari, N. Covalent Non-Fused Tetrathiafulvalene− Lyons, D. M.; Lynch, V. M.; Nielsen, K. A.; Jeppesen, J. O.; Kim, D.; et al.
Acceptor Systems. Chem. Commun. 2016, 52, 7906−7927. Ion-Regulated Allosteric Binding of Fullerenes (C60 and C70) by
(86) Bendikov, M.; Wudl, F.; Perepichka, D. F. Tetrathiafulvalenes, Tetrathiafulvalene-Calix[4]pyrroles. J. Am. Chem. Soc. 2014, 136,
Oligoacenenes, and Their Buckminsterfullerene Derivatives: The Brick 10410−10417.
and Mortar of Organic Electronics. Chem. Rev. 2004, 104, 4891−4945. (107) Nielsen, K. A.; Sarova, G. H.; Martín-Gomis, L.; Fernández-
(87) Geiser, U.; Schlueter, J. A. Conducting Organic Radical Cation Lázaro, F.; Stein, P. C.; Sanguinet, L.; Levillain, E.; Sessler, J. L.; Guldi,
Salts with Organic and Organometallic Anions. Chem. Rev. 2004, 104, D. M.; Sastre-Santos, Á .; Jeppesen, J. O. Chloride Anion Controlled
5203−5241. Molecular “Switching”. Binding of 2,5,7-Trinitro-9-dicyanomethylene-
(88) Kobayashi, H.; Cui, H. B.; Kobayashi, A. Organic Metals and fluorene-C60 by Tetrathiafulvalene Calix[4]pyrrole and Photophysical
Superconductors Based on BETS (BETS = bis(ethylenedithio)- Generation of Two Different Charge-Separated States. J. Am. Chem. Soc.
tetraselenafulvalene). Chem. Rev. 2004, 104, 5265−5288. 2008, 130, 460−462.
(89) Enoki, T.; Miyazaki, A. Magnetic TTF-Based Charge-Transfer (108) Nielsen, K. A.; Martín-Gomis, L.; Sarova, G. H.; Sanguinet, L.;
Complexes. Chem. Rev. 2004, 104, 5449−5477. Gross, D. E.; Fernández-Lázaro, F.; Stein, P. C.; Levillain, E.; Sessler, J.
(90) Ouahab, L. Coordination Complexes in Conducting and L.; Guldi, D. M.; et al. Binding Studies of Tetrathiafulvalene-
Magnetic Molecular Materials. Coord. Chem. Rev. 1998, 178−180, calix[4]pyrroles with Electron-Deficient Guests. Tetrahedron 2008, 64,
1501−1531. 8449−8463.
(91) Jeppesen, J. O.; Takimiya, K.; Jensen, F.; Brimert, T.; Nielsen, K.; (109) Jensen, L. G.; Nielsen, K. A.; Breton, T.; Sessler, J. L.; Jeppesen, J.
Thorup, N.; Becher, J. Pyrrolo-Annelated Tetrathiafulvalenes: The O.; Levillain, E.; Sanguinet, L. Self-Assembled Monolayers of Mono-
Parent Systems. J. Org. Chem. 2000, 65, 5794−5805. Tetrathiafulvalene Calix[4]pyrroles and Their Electrochemical Sensing
(92) Balandier, J.-Y.; Belyasmine, A.; Sallé, M. A Straightforward of Chloride. Chem. - Eur. J. 2009, 15, 8128−8133.
Access to Mono- and Bis(pyrrolo)tetrathiafulvalenes. Synthesis 2006, (110) Park, J. S.; Yoon, K. Y.; Kim, D. S.; Lynch, V. M.; Bielawski, C.
2006 (17), 2815−2817. W.; Johnston, K. P.; Sessler, J. L. Chemoresponsive Alternating

2706 DOI: 10.1021/acs.chemrev.6b00375


Chem. Rev. 2017, 117, 2641−2710
Chemical Reviews Review

Supramolecular Copolymers Created from Heterocomplementary (128) Medforth, C. J.; Senge, M. O.; Smith, K. M.; Sparks, L. D.;
Calix[4]pyrroles. Proc. Natl. Acad. Sci. U. S. A. 2011, 108, 20913−20917. Shelnutt, J. A. Nonplanar Distortion Modes for Highly Substituted
(111) Kim, D. S.; Lynch, V. M.; Park, J. S.; Sessler, J. L. Three Distinct Porphyrins. J. Am. Chem. Soc. 1992, 114, 9859−9869.
Equilibrium States via Self-Assembly: Simple Access to a Supra- (129) Nurco, D. J.; Medforth, C. J.; Forsyth, T. P.; Olmstead, M. M.;
molecular Ion-Controlled NAND Logic Gate. J. Am. Chem. Soc. 2013, Smith, K. M. Conformational Flexibility in Dodecasubstituted
135, 14889−14894. Porphyrins. J. Am. Chem. Soc. 1996, 118, 10918−10919.
(112) Bähring, S.; Martín-Gomis, L.; Olsen, G.; Nielsen, K. A.; Kim, D. (130) Sazanovich, I. V.; Galievsky, V. A.; van Hoek, A.; Schaafsma, T. J.;
S.; Duedal, T.; Sastre-Santos, Á .; Jeppesen, J. O.; Sessler, J. L. Design and Malinovskii, V. L.; Holten, D.; Chirvony, V. S. Photophysical and
Sensing Properties of a Self-Assembled Supramolecular Oligomer. Structural Properties of Saddle-Shaped Free-Base Porphyrins: Evidence
Chem. - Eur. J. 2016, 22, 1958−1967. for an “Orthogonal” Dipole Moment. J. Phys. Chem. B 2001, 105, 7818−
(113) Bähring, S.; Kim, D. S.; Duedal, T.; Lynch, V. M.; Nielsen, K. A.; 7829.
Jeppesen, J. O.; Sessler, J. L. Use of Solvent to Regulate the Degree of (131) Kadish, K. M.; Lin, M.; Caemelbecke, E. V.; De Stefano, G.;
Polymerisation in Weakly Associated Supramolecular Oligomers. Chem. Medforth, C. J.; Nurco, D. J.; Nelson, N. Y.; Krattinger, B.; Muzzi, C. M.;
Commun. 2014, 50, 5497−5499. Jaquinod, L.; et al. Influence of Electronic and Structural Effects on the
(114) Kim, D. S.; Chang, J.; Leem, D.; Park, J. S.; Thordarson, P.; Oxidative Behavior of Nickel Porphyrins. Inorg. Chem. 2002, 41, 6673−
Sessler, J. L. Redox- and pH-Responsive Orthogonal Supramolecular 6687.
Self-Assembly: An Ensemble Displaying Molecular Switching Charac- (132) Terazono, Y.; Dolphin, D. Synthesis and Characterization of β-
teristics. J. Am. Chem. Soc. 2015, 137, 16038−16042. Trifluoromethyl-meso-tetraphenylporphyrins. J. Org. Chem. 2003, 68,
(115) Park, J. S.; Karnas, E.; Ohkubo, K.; Chen, P.; Kadish, K. M.; 1892−1900.
Fukuzumi, S.; Bielawski, C. W.; Hudnall, T. W.; Lynch, V. M.; Sessler, J. (133) The Porphyrins Dolphin, D., Ed.; Academic Press: New York,
L. Ion-Mediated Electron Transfer in a Supramolecular Donor− 1978.
Acceptor Ensemble. Science 2010, 329, 1324−1327. (134) Milgrom, L. R. The Colours of Life: An Introduction to the
(116) Fukuzumi, S.; Ohkubo, K.; Kawashima, Y.; Kim, D. S.; Park, J. S.; Chemistry of Porphyrins and Related Compounds; Oxford University
Jana, A.; Lynch, V. M.; Kim, D.; Sessler, J. L. Ion-Controlled On-Off Press: Oxford, U.K., 1997.
Switching of Electron Transfer from Tetrathiafulvalene Calix[4]pyrroles (135) Electroanalytical Methods: Guide to Experiments and Applications;
to Li@C60. J. Am. Chem. Soc. 2011, 133, 15938−15941. Scholz, F., Ed.; Springer-Verlag: Berlin, 2010.
(117) Aoyagi, S.; Nishibori, E.; Sawa, H.; Sugimoto, K.; Takata, M.; (136) Jana, A.; Gobeze, H. B.; Ishida, M.; Mori, T.; Ariga, K.; Hill, J. P.;
Miyata, Y.; Kitaura, R.; Shinohara, H.; Okada, H.; Sakai, T.; et al. A D’Souza, F. Breaking Aggregation in a Tetrathiafulvalene-Fused Zinc
Layered Ionic Crystal of Polar Li@C60 Superatoms. Nat. Chem. 2010, 2, Porphyrin by Metal−Ligand Coordination to form a Donor−Acceptor
Hybrid for Ultrafast Charge Separation and Charge Stabilization. Dalton
678−683.
(118) Davis, C. M.; Kawashima, Y.; Ohkubo, K.; Lim, J. M.; Kim, D.; Trans. 2015, 44, 359−367.
(137) Wan, Z.; Jia, C.; Zhang, J.; Yao, X.; Shi, Y. Highly Conjugated
Fukuzumi, S.; Sessler, J. L. Photoinduced Electron Transfer from a
Donor−Acceptor Dyad Based on Tetrathiafulvalene Covalently
Tetrathiafulvalene-Calix[4]pyrrole to a Porphyrin Carboxylate within a
Attached to Porphyrin Unit. Dyes Pigm. 2012, 93, 1456−1462.
Supramolecular Ensemble. J. Phys. Chem. C 2014, 118, 13503−13513.
(138) Jia, H.; Schmid, B.; Liu, S.-X.; Jaggi, M.; Monbaron, P.; Bhosale,
(119) Davis, C. M.; Ohkubo, K.; Lammer, A. D.; Kim, D. S.;
S. V.; Rivadehi, S.; Langford, S. J.; Sanguinet, L.; Levillain, E.; et al.
Kawashima, Y.; Sessler, J. L.; Fukuzumi, S. Photoinduced Electron
Tetrathiafulvalene-Fused Porphyrins via Quinoxaline Linkers: Sym-
Transfer in a Supramolecular Triad Produced by Porphyrin Anion-
metric and Asymmetric Donor−Acceptor Systems. ChemPhysChem
induced Electron Transfer from Tetrathiafulvalene Calix[4]pyrrole to 2012, 13, 3370−3382.
Li+@C60. Chem. Commun. 2015, 51, 9789−9792. (139) Jana, A.; Ishida, M.; Cho, K.; Ghosh, S. K.; Kwak, K.; Ohkubo, K.;
(120) Bejger, C.; Davis, C. M.; Park, J. S.; Lynch, V. M.; Love, J. B.; Sung, Y. M.; Davis, C. M.; Lynch, V. M.; Lee, D.; et al.
Sessler, J. L. Palladium Induced Macrocyclic Preorganization for Tetrathiafulvalene-Annulated [28]Hexaphyrin(1.1.1.1.1.1): a Multi-
Stabilization of a Tetrathiafulvalene Mixed-Valence Dimer. Org. Lett. Electron Donor System Subject to Conformational Control. Chem.
2011, 13, 4902−4905. Commun. 2013, 49, 8937−8939.
(121) Kim, D.-S.; Lynch, V. M.; Nielsen, K. A.; Johnsen, C.; Jeppesen, J. (140) Jux, N. The Porphyrin Twist: Hückel and Möbius Aromaticity.
O.; Sessler, J. L. A Chloride-anion Insensitive Colorimetric Chemo- Angew. Chem., Int. Ed. 2008, 47, 2543−2546.
sensor for Trinitrobenzene and Picric Acid. Anal. Bioanal. Chem. 2009, (141) Osuka, A.; Saito, S. Expanded Porphyrins and Aromaticity.
395, 393−400. Chem. Commun. 2011, 47, 4330−4339.
(122) Poulsen, T.; Nielsen, K. A.; Bond, A. D.; Jeppesen, J. O. (142) Saito, S.; Osuka, A. Expanded Porphyrins: Intriguing Structures,
Bis(tetrathiafulvalene)-Calix[2]pyrrole[2]-Thiophene and Its Com- Electronic Properties, and Reactivities. Angew. Chem., Int. Ed. 2011, 50,
plexation with TCNQ. Org. Lett. 2007, 9, 5485−5488. 4342−4373 and references therein..
(123) Park, J. S.; Bejger, C.; Larsen, K. R.; Nielsen, K.; Jana, A.; Lynch, (143) Shin, J.-Y.; Kim, K. S.; Yoon, M.-C.; Lim, J. M.; Yoon, Z. S.;
V. M.; Jeppesen, J. O.; Kim, D.; Sessler, J. L. Synthesis and Recognition Osuka, A.; Kim, D. Aromaticity and Photophysical Properties of Various
Properties of Higher Order Tetrathiafulvalene (TTF) Calix[n]pyrroles Topology-Controlled Expanded Porphyrins. Chem. Soc. Rev. 2010, 39,
(n = 4−6). Chem. Sci. 2012, 3, 2685−2689. 2751−2767.
(124) Becher, J.; Brimert, T.; Jeppesen, J. O.; Pedersen, J. Z.; Zubarev, (144) Bolligarla, R.; Ishida, M.; Shetti, V. S.; Yamasumi, K.; Furuta, H.;
R.; Bjørnholm, T.; Reitzel, N.; Jensen, T. R.; Kjaer, K.; Levillain, E. Lee, C.-H. Intramolecular Charge Transfer Character in Tetrathiafulva-
Tetrathiafulvaleno-Annelated Porphyrins. Angew. Chem., Int. Ed. 2001, lene-Annulated Porphyrinoids: Effects of Core Modification and
40, 2497−2500. Protonation. Phys. Chem. Chem. Phys. 2015, 17, 8699−8705.
(125) Li, H.; Jeppesen, J. O.; Levillain, E.; Becher, J. A Mono-TTF- (145) Bill, N. L.; Ishida, M.; Bähring, S.; Lim, J. M.; Lee, S.; Davis, C.
Annulated Porphyrin as a Fluorescence Switch. Chem. Commun. 2003, M.; Lynch, V. M.; Nielsen, K. A.; Jeppesen, J. O.; Ohkubo, K.; et al.
846−847. Porphyrins Fused with Strongly Electron-Donating 1,3-Dithiol-2-
(126) Nielsen, K. A.; Levillain, E.; Lynch, V. M.; Sessler, J. L.; Jeppesen, ylidene Moieties: Redox Control by Metal Cation Complexation and
J. O. Tetrathiafulvalene Porphyrins. Chem. - Eur. J. 2009, 15, 506−516. Anion Binding. J. Am. Chem. Soc. 2013, 135, 10852−10862.
(127) Jana, A.; Ishida, M.; Kwak, K.; Sung, Y. M.; Kim, D. S.; Lynch, V. (146) Bill, N. L.; Ishida, M.; Kawashima, Y.; Ohkubo, K.; Sung, Y. M.;
M.; Lee, D.; Kim, D.; Sessler, J. L. Comparative Electrochemical and Lynch, V. M.; Lim, J. M.; Kim, D.; Sessler, J. L.; Fukuzumi, S. Long-Lived
Photophysical Studies of Tetrathiafulvalene-Annulated Porphyrins and Charge-Separated States Produced in Supramolecular Complexes
Their ZnII Complexes: The Effect of Metalation and Structural Between Anionic and Cationic Porphyrins. Chem. Sci. 2014, 5, 3888−
Variation. Chem. - Eur. J. 2013, 19, 338−349. 3896.

2707 DOI: 10.1021/acs.chemrev.6b00375


Chem. Rev. 2017, 117, 2641−2710
Chemical Reviews Review

(147) Liddell, P. A.; Kodis, G.; de la Garza, L.; Bahr, J. L.; Moore, A. L.; (166) Poddutoori, P. K.; Dion, A.; Yang, S.; Pilkington, M.; Wallis, J.
Moore, T. A.; Gust, D. Photoinduced Electron Transfer in D.; van der Est, A. Light-Induced Hole Transfer in a Hypervalent
Tetrathiafulvalene−Porphyrin−Fullerene Molecular Triads. Helv. Phosphorus(V) Octaethylporphyrin Bearing an Axially Linked bis-
Chim. Acta 2001, 84, 2765−2783. (ethylenedithio) Tetrathiafulvalene. J. Porphyrins Phthalocyanines 2010,
(148) Kodis, G.; Liddell, P. A.; de la Garza, L.; Moore, A. L.; Moore, T. 14, 178−187.
A.; Gust, D. Photoinduced Electron Transfer in π-Extended (167) Poddutoori, P. K.; Zarrabi, N.; Moiseev, A. G.; Gumbau-Brisa,
Tetrathiafulvalene−Porphyrin−Fullerene Triad Molecules. J. Mater. R.; Vassiliev, S.; van der Est, A. Long-Lived Charge Separation in Novel
Chem. 2002, 12, 2100−2108. Axial Donor−Porphyrin−Acceptor Triads Based on Tetrathiafulvalene,
(149) Di Valentin, M.; Bisol, A.; Agostini, G.; Liddell, P. A.; Kodis, G.; Aluminum(III) Porphyrin and Naphthalenediimide. Chem. - Eur. J.
Moore, A. L.; Moore, T. A.; Gust, D.; Carbonera, D. Photoinduced 2013, 19, 3148−3161.
Long-Lived Charge Separation in a Tetrathiafulvalene-Porphyrin- (168) van der Est, A.; Poddutoori, P. K. Light-Induced Spin
Fullerene Triad Detected by Time-Resolved Electron Paramagnetic Polarization in Porphyrin-Based Donor−Acceptor Dyads and Triads.
Resonance. J. Phys. Chem. B 2005, 109, 14401−14409. Appl. Magn. Reson. 2013, 44, 301−318.
(150) Sadaike, S.; Takimiya, K.; Aso, Y.; Otsubo, T. TTF−Porphyrin (169) Poddutoori, P. K.; Lim, G. N.; Sandanayaka, A. S. D.; Karr, P. A.;
Dyads as Novel Photoinduced Electron Transfer Systems. Tetrahedron Ito, O.; D’Souza, F.; Pilkington, M.; van der Est, A. Axially Assembled
Lett. 2003, 44, 161−165. Photosynthetic Reaction Center Mimics Composed of Tetrathiafulva-
(151) Liu, Y.; Wang, C.; Li, M.; Lv, S.; Lai, G.; Shen, Y. A New lene, Aluminum(III) Porphyrin and Fullerene Entities. Nanoscale 2015,
Fluorescence Molecular Switch Incorporating TTF and Tetraphenyl- 7, 12151−12165.
porphyrin Units. J. Porphyrins Phthalocyanines 2007, 11, 729−735. (170) Poddutoori, P. K.; Bregles, L. P.; Lim, G. N.; Boland, P.; Kerr, R.
(152) Li, M.; Huang, R.; Wu, C.; Zuo, H.; Lai, G.; Shen, Y. Synthesis G.; D’Souza, F. Modulation of Energy Transfer into Sequential Electron
and Properties of Tetrathiafulvalene-Porphyrin Assemblies. Front. Transfer upon Axial Coordination of Tetrathiafulvalene in an
Chem. Sci. Eng. 2011, 5, 422−428. Aluminum(III) Porphyrin−Free-Base Porphyrin Dyad. Inorg. Chem.
(153) Saha, S.; Johansson, L. E.; Flood, A. H.; Tseng, H.-R.; Zink, J. I.; 2015, 54, 8482−8494.
Stoddart, J. F. Powering a Supramolecular Machine with a Photoactive (171) Kandrashkin, Y. E.; Poddutoori, P. K.; van der Est, A. Electron
Molecular Triad. Small 2005, 1, 87−90. Transfer Pathways in a Tetrathiafulvalene-Aluminum(III) Porphyrin-
(154) Saha, S.; Johansson, E.; Flood, A. H.; Tseng, H.-R.; Zink, J. I.; Free-Base Porphyrin Triad Studied Using Electron Spin Polarization.
Stoddart, J. F. A Photoactive Molecular Triad as a Nanoscale Power Appl. Magn. Reson. 2016, 47, 511−526.
Supply for a Supramolecular Machine. Chem. - Eur. J. 2005, 11, 6846− (172) Poddutoori, P. K.; Lim, G. N.; Vassiliev, S.; D’Souza, F. Ultrafast
6858. Charge Separation and Charge Stabilization in Axially Linked
(155) Saha, S.; Flood, A. H.; Stoddart, J. F.; Impellizzeri, S.; Silvi, S.;
‘Tetrathiafulvalene−Aluminum(III) Porphyrin−Gold(III) Porphyrin’
Venturi, M.; Credi, A. A Redox-Driven Multicomponent Molecular
Reaction Center Mimics. Phys. Chem. Chem. Phys. 2015, 17, 26346−
Shuttle. J. Am. Chem. Soc. 2007, 129, 12159−12171.
26358.
(156) Drobizhev, M.; Stepanenko, Y.; Dzenis, Y.; Karotki, A.; Rebane,
(173) Nakanishi, T.; Kojima, T.; Ohkubo, K.; Hasobe, T.; Nakayama,
A.; Taylor, P. N.; Anderson, H. L. Understanding Strong Two-Photon
K.; Fukuzumi, S. Photoconductivity of Porphyrin Nanochannels
Absorption in π-Conjugated Porphyrin Dimers via Double-Resonance
Composed of Diprotonated Porphyrin Dications with Saddle Distortion
Enhancement in a Three-Level Model. J. Am. Chem. Soc. 2004, 126,
15352−15353. and Electron Donors. Chem. Mater. 2008, 20, 7492−7500.
(157) Aratani, N.; Kim, D.; Osuka, A. π-Conjugation Enlargement (174) Honda, T.; Nakanishi, T.; Ohkubo, K.; Kojima, T.; Fukuzumi, S.
Toward the Creation of Multi-Porphyrinic Systems with Large Two- Structure and Photoinduced Electron Transfer Dynamics of a Series of
Photon Absorption Properties. Chem. - Asian J. 2009, 4, 1172−1182. Hydrogen-Bonded Supramolecular Complexes Composed of Electron
(158) Pawlicki, M.; Collins, H. A.; Denning, R. G.; Anderson, H. L. Donors and a Saddle-Distorted Diprotonated Porphyrin. J. Am. Chem.
Two-Photon Absorption and the Design of Two-Photon Dyes. Angew. Soc. 2010, 132, 10155−10163.
Chem., Int. Ed. 2009, 48, 3244−3266. (175) Zhang, Y.-M.; Chen, Y.; Zhuang, R.-J.; Liu, Y. Supramolecular
(159) Ogawa, K.; Nagatsuka, Y. Bisporphyrin Connected by Architecture of Tetrathiafulvalene-Bridged bis(β-cyclodextrin) with
Tetrathiafulvalene. J. Porphyrins Phthalocyanines 2009, 13, 114−121. Porphyrin and its Electron Transfer Behaviors. Photochem. Photobiol. Sci.
(160) Ogawa, K.; Dy, J.; Kobuke, Y. Substituent Effect on Two-Photon 2011, 10, 1393−1398.
Absorption Properties of Conjugated Porphyrins. J. Porphyrins (176) Chen, Y.; Liu, Y. Cyclodextrin-Based Bioactive Supramolecular
Phthalocyanines 2005, 9, 735−744. Assemblies. Chem. Soc. Rev. 2010, 39, 495−505.
(161) Liu, C.-G.; Guan, W.; Song, P.; Yan, L.-K.; Su, Z.-M. Redox- (177) Xue, L.-J.; Huo, P.; Li, Y.-H.; Hou, J.-L.; Zhu, Q.-Y.; Dai, J. An
Switchable Second-Order Nonlinear Optical Responses of Push−Pull Ionic Charge-Transfer Dyad Prepared Cost-Effectively from a
Monotetrathiafulvalene-Metalloporphyrins. Inorg. Chem. 2009, 48, Tetrathiafulvalene Carboxylate Anion and a TMPyP Cation. Phys.
6548−6554. Chem. Chem. Phys. 2016, 18, 2940−2948.
(162) Xiao, X.; Xu, W.; Zhang, D.; Xu, H.; Lu, H.; Zhu, D. A New (178) Kopranenkov, V. N.; Luk'yanets, E. A. Porphyrazines: Synthesis,
Fluorescence-Switch Based on Supermolecular Dyad with Properties, Application. Russ. Chem. Bull. 1995, 44, 2216−2232.
(Tetraphenylporphyrinato)Zinc(II) and Tetrathiafulvalene units. J. (179) Baerends, E. J.; Ricciardi, G.; Rosa, A.; van Gisbergen, S. J. A. A
Mater. Chem. 2005, 15, 2557−2561. DFT/TDDFT Interpretation of the Ground and Excited States of
(163) Boixel, J.; Fortage, J.; Blart, E.; Pellegrin, Y.; Hammarström, L.; Porphyrin and Porphyrazine Complexes. Coord. Chem. Rev. 2002, 230,
Becker, H.-C.; Odobel, F. Extension of the Charge Separated-State 5−27.
Lifetime by Supramolecular Association of a Tetrathiafulvalene Electron (180) Chen, T.; Wang, C.; Qiu, H.; Jin, L.; Yin, B.; Imafuku, K.
Donor to a Zinc/Gold Bisporphyrin. Dalton Trans. 2010, 39, 1450− Synthesis and Aggregation of Tetrathiafulvalene-Annulated Porphyr-
1452. azines. Heterocycles 2007, 71, 549−555.
(164) Fortage, J.; Boixel, J.; Blart, E.; Hammarström, L.; Becker, H. C.; (181) Hou, R.; Jin, L.; Yin, B. Synthesis and Electron Donating
Odobel, F. Single-Step Electron Transfer on the Nanometer Scale: Property of Novel Porphyrazines Containing Tetrathiacrown Ether-
Ultra-Fast Charge Shift in Strongly Coupled Zinc Porphyrin−Gold Linked Tetrathiafulvalene Moieties. Inorg. Chem. Commun. 2009, 12,
Porphyrin Dyads. Chem. - Eur. J. 2008, 14, 3467−3480. 739−743.
(165) Fortage, J.; Boixel, J.; Blart, E.; Becker, H. C.; Odobel, F. Very (182) Leng, F.; Hou, R.; Jin, L.; Yin, B.; Xiong, R.-G. Synthesis,
Fast Single-Step Photoinduced Charge Separation in Zinc Porphyrin Characterization and Electrochemistry of the Novel Metalloporphyr-
Bridged to a Gold Porphyrin by a Bisethynyl Quaterthiophene. Inorg. azines Annulated with Tetrathiafulvalene Having Pentoxycarbonyl
Chem. 2009, 48, 518−526. Substituents. J. Porphyrins Phthalocyanines 2010, 14, 108−114.

2708 DOI: 10.1021/acs.chemrev.6b00375


Chem. Rev. 2017, 117, 2641−2710
Chemical Reviews Review

(183) Leng, F.; Wang, X.; Jin, L.; Yin, B. The Synthesis and Properties (202) Sly, J.; Kasák, P.; Gomar-Nadal, E.; Rovira, C.; Górriz, L.;
of Unsymmetrical Porphyrazines Annulated with a Tetrathiafulvalene Thordarson, P.; Amabilino, D. B.; Rowan, A. E.; Nolte, R. J. M. Chiral
Bearing Two Tetraethylene Glycol Units. Dyes Pigm. 2010, 87, 89−94. Molecular Tapes from Novel Tetra(thiafulvalene-crown-ether)-Sub-
(184) Tao, W.; Kan, Y.-H.; Wu, S.-X.; Li, H.-B.; Yan, L.-K.; Sun, S.-L.; stituted Phthalocyanine Building Blocks. Chem. Commun. 2005, 1255−
Su, Z.-M. The Origin of the Unusual Broad and Intense Visible 1257.
Absorption of Tetrathiafulvalene-Annulated Zinc Porphyrazine: A (203) Elemans, J. A. A. W.; Van Hameren, R.; Nolte, R. J. M.; Rowan,
Density Functional Theory Study. J. Mol. Graphics Modell. 2012, 33, A. E. Molecular Materials by Self-Assembly of Porphyrins, Phthalocya-
26−34. nines, and Perylenes. Adv. Mater. 2006, 18, 1251−1266.
(185) Hou, R.; Jiang, C.; Yin, B. A Novel Tetra(tetrathiafulvalene- (204) Kimura, M.; Otsuji, S.; Takizawa, J.; Tatewaki, Y.; Fukawa, T.;
Thiacrown Ether)-Substituted Porphyrazine: Synthesis, Electrochem- Shirai, H. Supramolecular Stacks of Asymmetric Zinc Phthalocyanines
ical, and Control Aggregation in Solution by Complexation of Functionalized with One Tetrathiafulvalene Unit. Chem. Lett. 2010, 39,
Transition-Metal Ions. Heterocycles 2010, 81, 717−725. 812−813.
(186) Wudl, F.; Smith, G. M.; Hufnagel, E. J. Bis-1,3-Dithiolium (205) Kimura, T.; Watanabe, D.; Namauo, T. Preparation and Optical
Chloride: An Unusually Stable Organic Radical Cation. J. Chem. Soc. D and Electrochemical Properties of Octaethylphthalocyanines Fused with
1970, 21, 1453−1454. Several Tetrathiafulvalene Units. Heteroat. Chem. 2011, 22, 605−611.
(187) Nygaard, S.; Laursen, B. W.; Hansen, T. S.; Bond, A. D.; Flood, (206) Kimura, T.; Takahashi, N.; Tajima, T.; Takaguchi, Y. Preparation
A. H.; Jeppesen, J. O. Preparation of Cyclobis(paraquat-p-phenylene)- and Optical and Electrochemical Properties of Unsymmetrical
Based [2]Rotaxanes Without Flexible Glycol Chains. Angew. Chem. Phthalocyanines with One or Two TTF Units. Heterocycles 2012, 84,
2007, 119, 6205−6209. 333−337.
(188) Akutagawa, T.; Kakiuchi, K.; Hasegawa, T.; Nakamura, T.; (207) Kimura, T.; Takahashi, N.; Tajima, T.; Takaguchi, Y. α-
Christensen, C. A.; Becher, J. Langmuir−Blodgett Films of Amphiphilic Substituted Unsymmetrical Phthanlocyanines with One through Three
Bis(tetrathiafulvalene) Macrocycles with Four Alkyl Chains. Langmuir TTF units and Their Optical and Electrochemical Properties.
2004, 20, 4187−4195. Heterocycles 2012, 86, 679−686.
(189) Jain, A.; Rao, K. V.; Mogera, U.; Sagade, A. A.; George, S. J. (208) Cook, M. J.; Cooke, G.; Jafari-Fini, A. A Liquid Crystalline
Dynamic Self-Assembly of Charge-Transfer Nanofibers of Tetrathia- Tetrathiafulvalenylphthalocyanine. Chem. Commun. 1996, 1925−1926.
fulvalene Derivatives with F4TCNQ. Chem. - Eur. J. 2011, 17, 12355− (209) Blower, M. A.; Bryce, M. R.; Devonport, W. Synthesis and
12361. Aggregation of a Phthalocyanine Symmetrically-Functionalized with
(190) Hou, R.; Qiu, H.; Chen, T.; Yin, B. Synthesis, Photophysical and Eight Tetrathiafulvalene Units. Adv. Mater. 1996, 8, 63−65.
Electrochemical Properties of a D−σ−A Ensemble Derived from (210) Wang, C. S.; Bryce, M. R.; Batsanov, A. S.; Stanley, C. F.; Beeby,
Porphyrazine and Tetrathiafulvalene. Heterocycles 2009, 78, 1799−1805. A.; Howard, J. A. K. Synthesis, Spectroscopy and Electrochemistry of
(191) Wang, C.; Bryce, M. R.; Batsanov, A. S.; Howard, J. A. K. Phthalocyanine Derivatives Functionalised with Four and Eight
Synthesis of Pyrazinoporphyrazine Derivatives Functionalised with
Peripheral Tetrathiafulvalene Units. J. Chem. Soc., Perkin Trans. 2
Tetrathiafulvaiene (TTF) Units: X-Ray Crystal Structures of Two
1997, 1671−1678.
Related TTF Cyclophanes and Two Bis(1,3-dithiole-2-thione)
(211) Hu, Y.; Shen, Y. Synthesis of Zinc Phthalocyanine Derivative
Intermediates. Chem. - Eur. J. 1997, 3, 1679−1690.
Functionalized with Four Peripheral Tetrathiafulvalene Units. J.
(192) Leznoff, C. C., Lever, A. B. P., Eds. Phthalocyanines, Properties and
Heterocycl. Chem. 2002, 39, 1071−1075.
Applications; Wiley-VCH: New York, 1989−1996; Vols. 1−4.
(212) Hu, Y.; Lai, G.; Shen, Y.; Li, Y. Synthesis, Optical Spectroscopy
(193) Stuzhin, P. A.; Ercolani, C. Porphyrazines with Annulated
and Electrochemistry of a D−σ−A Compound Derived from
Heterocycles. In The Porphyrin Handbook; Kadish, K. M., Smith, K. M.,
Magnesium Phthalocyanine. J. Porphyrins Phthalocyanines 2004, 8,
Guilard, R., Eds.; Academic Press: New York, 2003; Vol. 15, pp 263−
264. 1042−1046.
(194) Dolphin, D. The Porphyrins; Academic Press: New York, 1978− (213) Hu, Y.; Lai, G.; Shen, Y.; Li, Y. Synthesis, Spectroscopy, and
1979; Vols. 1−7. Electrochemistry of Metallophthalocyanines Substituted by Propylene-
(195) Loosli, C.; Jia, C.; Liu, S.-X.; Haas, M.; Dias, M.; Levillain, E.; dithiotetrathiafulvalene Derivatives. Monatsh. Chem. 2004, 135, 1167−
Neels, A.; Labat, G.; Hauser, A.; Decurtins, S. Synthesis and 1172.
Electrochemical and Photophysical Studies of Tetrathiafulvalene- (214) Hu, Y. Y.; Lai, G. Q.; Shen, Y. J.; Li, Y. F. Synthesis, Optical
Annulated Phthalocyanines. J. Org. Chem. 2005, 70, 4988−4992. Spectroscopy and Electrochemistry of TTF-Derived Metallophthalo-
(196) Donders, C. A.; Liu, S.-X.; Loosli, C.; Sanguinet, L.; Neels, A.; cyanine Complexes. Dyes Pigm. 2005, 66, 49−53.
Decurtins, S. Synthesis of Tetrathiafulvalene-Annulated Phthalocya- (215) Farren, C.; Christensen, C. A.; FitzGerald, S.; Bryce, M. R.;
nines. Tetrahedron 2006, 62, 3543−3549. Beeby, A. Synthesis of Novel Phthalocyanine−Tetrathiafulvalene
(197) Kimura, T.; Watanabe, D.; Namauo, T. Preparation and Hybrids; Intramolecular Fluorescence Quenching Related to Molecular
Electrochemical Property of Octabutylphthalocyanine Fused with Four Geometry. J. Org. Chem. 2002, 67, 9130−9139.
TTF Units. Heterocycles 2008, 76, 1023−1026. (216) Baumann, T. F.; Nasir, M. S.; Sibert, J. W.; White, A. J. P.;
(198) Kimura, T.; Namauo, T.; Amano, K.; Takahashi, N.; Takaguchi, Olmstead, M. M.; Williams, D. J.; Barrett, A. G. M.; Hoffman, B. M.
Y.; Hoshi, T.; Kobayashi, N. Preparation and Electrochemical and Solitaire-Porphyrazines: Synthetic, Structural, and Spectroscopic
Optical Properties of α-Octaalkylphthalocyanines with Four Fused TTF Investigation of Complexes of the Novel Binucleating Norphthalocya-
Units. J. Porphyrins Phthalocyanines 2011, 15, 547−554. nine-2,3-dithiolato Ligand. J. Am. Chem. Soc. 1996, 118, 10479−10486.
(199) Kimura, T.; Takahashi, N.; Tajima, T.; Takaguchi, Y. Preparation (217) Beall, L. S.; Mani, N. S.; White, A. J.; Williams, D. J.; Barrett, A.
and Optical and Electrochemical Properties of Phthalocyanines with the G.; Hoffman, B. M. Porphyrazines and Norphthalocyanines Bearing
TTF Unit. Phosphorus, Sulfur Silicon Relat. Elem. 2013, 188, 408−412. Nitrogen Donor Pockets: Metal Sensor Properties. J. Org. Chem. 1998,
(200) Kimura, T.; Chiba, T.; Nakajo, S. Preparation of Phthalocyanines 63, 5806−5817.
Fused with Four Tetrathiafulvalene Units via the Functional Group (218) Rodríguez-Morgade, M. S.; de la Torre, G.; Torres, T. Design
Transformation of Tetrakis(o-xylylenedithio)phthalocyanines with and Synthesis of Low-Symmetry Phthalocyanines and Related Systems.
Toluene/Aluminum Chloride. J. Porphyrins Phthalocyanines 2014, 18, In The Porphyrin Handbook; Kadish, K. M., Smith, K. M., Guilard, R.,
884−890. Eds.; Academic Press: New York, 2003; Vol. 15, pp 125−160.
(201) Wang, R.; Liu, W.; Chen, Y.; Zuo, J.-L.; You, X.-Z. The Synthesis (219) Guo, J.; Xia, Y.; Li, D.; Hou, R. Synthesis and Electron-Donating
and Electrochemical Properties of a New Tetra-(crown-ether- Properties of Novel Norphthalocyanines Containing Thiacrown Ether-
thiafulvalene)-Annulated Phthalocyanine Derivative. Dyes Pigm. 2009, Linked Tetrathiafulvalene Moieties. Tetrahedron Lett. 2016, 57, 570−
81, 40−44. 573.

2709 DOI: 10.1021/acs.chemrev.6b00375


Chem. Rev. 2017, 117, 2641−2710
Chemical Reviews Review

(220) Hou, R.; Jiang, C.; Chen, T.; Jin, L.-Y.; Yin, B. The
Norphthalocyanines Bearing Two TTF Units: Synthesis, Photophysical
and Electrochemical Properties. Heterocycles 2011, 83, 1859−1866.
(221) Claessens, C. G.; González-Rodríguez, D.; Rodríguez-Morgade,
M. S.; Medina, A.; Torres, T. Subphthalocyanines, Subporphyrazines,
and Subporphyrins: Singular Nonplanar Aromatic Systems. Chem. Rev.
2014, 114, 2192−2277.
(222) Shimizu, S.; Yamazaki, Y.; Kobayashi, N. Tetrathiafulvalene-
Annulated Subphthalocyanines. Chem. - Eur. J. 2013, 19, 7324−7327.

2710 DOI: 10.1021/acs.chemrev.6b00375


Chem. Rev. 2017, 117, 2641−2710

You might also like