You are on page 1of 56

Review

pubs.acs.org/CR

Control and Switching of Aromaticity in Various All-Aza-Expanded


Porphyrins: Spectroscopic and Theoretical Analyses
Young Mo Sung,† Juwon Oh,† Won-Young Cha,† Woojae Kim,† Jong Min Lim,†,‡ Min-Chul Yoon,†,§
and Dongho Kim*,†

Spectroscopy Laboratory for Functional π-Electronic Systems and Department of Chemistry, Yonsei University, Seoul 120-749,
South Korea

Physical and Theoretical Chemistry Laboratory, Department of Chemistry, University of Oxford, South Parks Road, Oxford OX1
3QZ, United Kingdom
§
Manufacturing Engineering Team, Memory Manufacturing Operation Center, Samsung Electronics, Samsungjeonja-ro 1,
Hwasung-si, Gyeonggi-do 18448, South Korea

ABSTRACT: Modification of aromaticity is regarded as one of the most interesting and


important research topics in the field of physical organic chemistry. Particularly,
porphyrins and their analogues (porphyrinoids) are attractive molecules for exploring
various types of aromaticity because most porphyrinoids exhibit circular conjugation
pathways in their macrocyclic rings with various molecular structures. Aromaticity in
porphyrinoids is significantly affected by structural modification, redox chemistry, NH
tautomerization, and electronic states (singlet and triplet excited states). Conversely,
aromaticity significantly affects the spectroscopic properties and chemical reactivities of
porphyrinoids. In this context, considerable efforts have been devoted to understanding
and controlling the aromaticity and antiaromaticity of porphyrinoids. Thus, a series of
porphyrinoids are in the limelight, being expected to shed light on this field because they
have some advantages to demonstrate the switching of aromaticity; it is possible to
control the aromaticity by lowering the temperature, adding and removing the protons
of expanded porphyrins, changing the chemical environment, and switching the
electronic states (triplet and singlet excited states) by photoexcitation. In this regard, this Review describes the control of
aromaticity in various expanded porphyrins from the spectroscopic point of view with assistance from theoretical calculations.

CONTENTS 3.2.3. Deprotonation Effects in Larger Ex-


panded Porphyrins 2289
1. Introduction 2258 4. Control of Aromaticity by Various Solvents 2291
2. Control of Aromaticity by Regulating the Temper- 4.1. Control of Aromaticity in Heptaphyrin by
ature 2260 Solvent 2292
2.1. Change of Aromaticity in meso-Hexakis- 5. Reversal of Aromaticity in the Excited State 2297
(pentafluorophenyl) Hexaphyrin at Low 5.1. Reversal of Aromaticity in the Lowest Triplet
Temperatures 2260 State 2298
2.2. Change of Aromaticity in meso-Hexakis- 5.2. Reversal of Aromaticity in the Singlet Excited
(trifluoromethyl) Hexaphyrins at Low Tem- State 2301
peratures 2265 5.3. Reversal of Aromaticity for Mö bius Com-
3. Control of Aromaticity by Protonation and pounds in the Triplet Excited State 2303
Deprotonation 2272 6. Summary and Outlook 2305
3.1. Control of Aromaticity by Protonation 2272 Author Information 2307
3.1.1. Effects of Protonation in [4n+2]π Ex- Corresponding Author 2307
panded Porphyrins 2272 ORCID 2307
3.1.2. Effects of Protonation in [4n]π Expanded Notes 2307
Porphyrins 2277 Biographies 2307
3.2. Control of Aromaticity by Deprotonation 2283 Acknowledgments 2307
3.2.1. Effects of Deprotonation in [4n+2]π
Expanded Porphyrins 2283
3.2.2. Effects of Deprotonation in [4n]π Ex- Special Issue: Expanded, Contracted, and Isomeric Porphyrins
panded Porphyrins 2287
Received: May 16, 2016
Published: December 16, 2016

© 2016 American Chemical Society 2257 DOI: 10.1021/acs.chemrev.6b00313


Chem. Rev. 2017, 117, 2257−2312
Chemical Reviews Review

References 2308 theory proposed by Hückel, various concepts of aromaticity have


been proposed, including half-twisted Möbius, doubly twisted
(figure-eight structures), and three-dimensional (spherical)
1. INTRODUCTION conjugation compounds.41,42 In 1964, Heilbronner reported
Expanded porphyrins, which are macrocyclic compounds that stable molecular systems with [4n]π electrons, exhibiting
containing more than four pyrrole or pyrrole-related rings closed-shell configuration, can be achieved by distorting the
(e.g., furan, thiophene, and/or other heterocyclic subunits), have molecular structures to a Möbius strip topology.45 In his report,
attracted significant attention over the past few decades.1−8 Since compounds with Möbius structures exhibit aromatic and
the first report of expanded porphyrin named “sapphyrin” in antiaromatic character in [4n] and [4n+2]π electronic systems,
1966 by Woodward and co-workers,9,10 numerous studies about respectively, which is in direct contrast to the prediction made by
expanded porphyrins have been reported.11 In 1975, Day, Marks, the Hückel [4n+2] rule. Typically, nonplanar steric multi-
and Wachter reported a uranyl−superphthalocyanine complex, conjugated annulenes can be divided into Hückel or Möbius
representing one of the most important characterization studies molecules on the basis of their structural topologies by
in the synthesis of expanded porphyrin.12 After that, in 1975, theoretical methods employing a coiled ribbon model proposed
Berger and LeGoff reported other significant examples of by Rzepa.42 In their expanded definition, twisted molecules with
expanded porphyrins, such as [22]platyrin; in 1985, Rexhausen an even integer for the linking number are aromatic when they
and Gossauer reported [22]pentaphyrin.13,14 In the 1990s, have [4n+2]π electrons that satisfy the Hückel [4n+2] rule. By
Sessler et al. reported the synthesis of expanded porphyrins, such contrast, entangled molecules with an odd integer for the linking
as [22]sapphyrin, [24]amethyrin, [24]rosarin, [26]rubyrin, number are aromatic when they have [4n]π electrons, referred to
[28]heptaphyrin, and [40]turcasarin.15−17 In addition, they as Möbius aromaticity.
reviewed various synthetic procedures and demonstrated the In company with the concepts of aromaticity based on
potential applications of expanded porphyrins for applications, molecular topology, in 1972, Baird proposed the modulation of
such as photodynamic therapy (PDT) and anion recognition, as electronic features of the lowest triplet state.46 More specifically,
well as their use in functional near-infrared (NIR) dyes for Baird predicted that the aromaticity of planar conjugated ring
nonlinear optical (NLO) applications.15,16,18−29 Furthermore, systems in the ground state is reversed in their lowest triplet state,
Chandrashekar and co-workers have synthesized several core- which was developed by utilizing the perturbation molecular
modified expanded porphyrins by substituting pyrroles with orbital theory. Similar to the concept of aromaticity in Möbius
furans, thiophenes, and selenophenes and investigated their topology, the electron-counting rule for defining aromaticity in
structural diversity and degree of aromaticity.30−34 In particular, the lowest triplet state is contrary to the Hückel [4n+2] rule;
in 2007, the Latos-Grażyński group successfully synthesized di-p- annulenes having [4n]π electrons would be aromatic, while
benzi[28]hexaphyrin(1.1.1.1.1.1), which exhibits an interesting annulenes with [4n+2]π electrons would be antiaromatic. Baird
phenomenon: Hückel−Möbius aromaticity switching depending demonstrated the reversal of aromaticity in the lowest triplet
on temperature.35 In addition, the Latos-Grażyński group has state, while a number of theoretical studies have been conducted
revealed the relationship between molecular conformation and to extend Baird’s rule to the first singlet excited state.47,48 In 2008,
aromaticity, which was mainly characterized by NMR spectros- Karadakov suggested that the aromaticity of annulene in the
copy and quantum chemical calculations.36 Since this intriguing lowest singlet excited state is also possibly reversed as compared
finding of Hückel−Möbius aromaticity switching, the groups of with that in its ground state.47,48
Kim and Osuka have extended the field of Möbius expanded For confirming these various concepts of aromaticity, a
porphyrins by synthesizing and characterizing various Möbius significant number of methods for evaluating the degree of
(anti)aromatic expanded porphyrins37−39 and, by extension, aromaticity have been investigated, especially for quantum
Hückel (anti)aromatic expanded porphyrins. Although all mechanical calculations. The calculated indices for aromaticity
expanded porphyrins are tremendously interesting, this Review are divided into three groups: energetic aromaticity indices,
focuses on all-aza-expanded porphyrins. topological aromaticity indices, and magnetic-based aromaticity
Among various subjects in the study of expanded porphyrins, indices. Energetic aromaticity indices are represented as aromatic
aromaticity is the topic of interest, which has been explored from stabilization energy (ASE), indicative of the energy difference
both theoretical and experimental perspectives.11,37−39 Aroma- between aromatic (antiaromatic) and nonaromatic isomers.49,50
ticity, which is a measure of the unique stability of conjugated Topological aromaticity indices, such as bond length alternation
ring systems, has been regarded as one of the most intriguing and (BLA) and harmonic oscillator model of aromaticity (HOMA),
significant concepts in chemistry, dating back to the discovery of are considerably utilized and also have been regarded as
benzene by Faraday in 1825 and its proposed resonance structure approximate indicators of aromaticity.41,51 Magnetic aromaticity
by Kekulé in 1865.40 Numerous studies have been reported on indices, represented by nucleus-independent chemical shift
aromaticity as the degree of aromaticity in conjugated ring (NICS), are considered to be precise and powerful for estimating
systems determines their chemical reactivity and properties.41−43 the degree of aromaticity.52
Furthermore, aromaticity can be utilized for various applications, With these theoretical aromaticity indices, experimental
such as the synthesis of novel π-conjugated systems and aromaticity indices have also been in focus. As antiaromatic
development of synthetic routes via the prediction of reactant annulenes are energetically unstable, it has been difficult to
stability. In 1931, Hückel reported the [4n+2] rule for conjugated synthesize stable antiaromatic annulenes, thereby impeding
ring systems.44 In his theory, annulenes having [4n+2]π studies on the experimental aromaticity indices. Nevertheless,
electrons, such as benzene, exhibit closed electronic config- since stable aromatic and antiaromatic expanded porphyrins
uration, demonstrating aromatic character, while annulenes having more than four pyrroles with a large cavity size have been
having [4n]π electrons, such as cyclobutadiene, exhibit open- successfully synthesized,11,38,53 various physical properties, such
shell configuration with electrons in nonbonding molecular as energetic, magnetic, structural, and photophysical properties,
orbitals (MOs), demonstrating antiaromatic character. Since the have been suggested to be utilized for confirming whether the
2258 DOI: 10.1021/acs.chemrev.6b00313
Chem. Rev. 2017, 117, 2257−2312
Chemical Reviews Review

molecule is aromatic, antiaromatic, or nonaromatic.40 Accord-


ingly, expanded porphyrins are now regarded as the optimum
test subjects for investigating the correlation between molecular
properties and (anti)aromaticity and exploring novel aromatic
concepts for the experimental evaluation of molecular (anti)-
aromaticity.
As part of our efforts toward experimental aromaticity indices,
in-depth analyses have demonstrated that various photophysical
properties, such as excited-state lifetime, absorption spectral
features, and fluorescence, can be utilized for confirming whether
expanded porphyrins are aromatic, antiaromatic, or non-
aromatic.38,39,54 In particular, expanded porphyrins exhibit
unique absorption and emission spectral features depending on
their aromaticity. For aromatic expanded porphyrins, intense and
sharp B-like (also called Soret-like band) and weak but structured
Q-like bands are characteristically observed around visible and
NIR regions, respectively, in their absorption spectra; these
spectral features can be explained on the basis of Gouterman’s
four-orbital model.55 Aromatic expanded porphyrins exhibit two
pairs of frontier molecular orbitals (FMOs), almost degenerate,
where the configuration interaction between these four FMOs
results in intense B-like and weak Q-like absorption bands. In
addition, in aromatic expanded porphyrins, fluorescence in the
NIR region is characterized by vibronic bands.38 On the other
hand, antiaromatic expanded porphyrins exhibit different
features. As compared to aromatic ones, antiaromatic expanded
porphyrins exhibit significantly attenuated absorption bands in
the visible region with characteristic absorption tailing over the Figure 1. Schematic diagrams for molecular orbitals and electronic
NIR region.54 Moreover, antiaromatic expanded porphyrins do structures of aromatic and antiaromatic expanded porphyrins.
not exhibit fluorescence. These absorption and emission spectral
features are understood by their MO structures and energy intense GSB signals with relatively weak ESA signals on both
diagram. In contrast to the four degenerate FMOs of aromatic sides of the GSB signal in the visible region; they also exhibit long
expanded porphyrins, antiaromatic ones exhibit nondegenerate excited-state lifetimes, which is in good agreement with their
highest occupied molecular orbital (HOMO) and lowest fluorescent property.38 Notably, the excited-state lifetime is
unoccupied molecular orbital (LUMO) with a small energy significantly shorter than those of H2TPP (10 ns) and ZnTPP (2
gap (Figure 1). Because these HOMOs and LUMOs have gerade ns), attributed to the flexible structures of expanded porphyrins
symmetry, the lowest electronic transition between HOMO and in addition to the reduced HOMO−LUMO energy gap.
LUMO is optically forbidden by the selection rule, resulting in Expanded porphyrins with rigid structures formed by internal
the tailing absorption in the NIR region.39 In addition, the energy bridges and fused constituents exhibit a large increase in the S1
gaps between HOMO−1/HOMO−2 and LUMO+1/LUMO+2 state lifetime.54,57,58 On the other hand, significantly different TA
are decreased as compared with those of aromatic expanded spectral features and excited-state dynamics are observed in
porphyrins, in which HOMO → LUMO+1/LUMO+2 and antiaromatic expanded porphyrins. Their TA spectra exhibit low-
HOMO−1/HOMO−2 → LUMO transitions mainly result in intensity GSB signals and strong ESA signals in the visible region,
relatively weak absorption bands in the visible region. which is in sharp contrast to the high-intensity GSB and weaker
Interestingly, these different features of expanded porphyrins, ESA signals observed in the TA spectra of aromatic expanded
which are dependent on their aromaticity, are extended to the porphyrins.39 Furthermore, although antiaromatic expanded
results obtained from transient absorption (TA) measurements. porphyrins are rapidly relaxed from the excited state to the
TA spectroscopy measures a difference absorption (ΔA) ground state, they exhibit distinctive excited-state dynamics, in
spectrum (absorption spectrum of the species in the excited which two-stepwise relaxation processes (one in the order of
state minus absorption spectrum of the species in the ground hundreds of femtoseconds and the other in the order of tens of
state) via the promotion of molecules to an electronically excited picoseconds) are observed.38 Excited-state dynamics is distin-
state for typically investigating their excited-state dynamics guished from typical structural relaxation in the lowest excited
(Figure 2).56 As molecules (maximum up to 10%) can be excited state because structurally locked expanded porphyrins also
by a pump pulse, the TA spectra comprise the following three exhibit similar relaxation dynamics without significant
main features: (1) ground-state bleaching (GSB) signals, changes.54,57,58 This unique excited-state dynamics is in
representing negative signals in ΔA attributed to the decrease accordance with the presence of the optically dark state, as
of ground-state absorption, caused by vacancies in the ground- indicated by absorption tailing in the NIR region; the optically
state population by the excitation of molecules; (2) stimulated dark state facilitates and accelerates two-stepwise internal
emission (SE) signals, observed when the excited population conversion (relaxation) processes, which also indicate that the
returns to the ground state after the passage of the probe pulse; excited-state dynamics of antiaromatic expanded porphyrins are
and (3) excited-state absorption (ESA) signals, attributed to less sensitive to the conformational effect.
absorption by excited molecules to high energy levels. For In addition, aromaticity-dependent behaviors could be
aromatic expanded porphyrins, their intense B-like bands exhibit observed in a two-photon absorption (TPA) phenomenon,
2259 DOI: 10.1021/acs.chemrev.6b00313
Chem. Rev. 2017, 117, 2257−2312
Chemical Reviews Review

Figure 2. Schematic illustration of TA spectroscopy (right) and spectral components to TA spectrum (left).

Chart 1. Schematic Molecular Structures of PF26H and PF28H (Reprinted with Permission from Ref 63; Copyright 2009
American Chemical Society)

which depends on second-order hyperpolarizability. The TPA involved change of molecular properties for experimentally
cross section value can be regarded as another important determining aromaticity.
experimental factor to determine the extent of aromaticity for In this Review, the control and switching of aromaticity in
expanded porphyrinoids. Notably, previous studies on a series of expanded porphyrins via their photophysical characterization
octupolar systems possessing electron donor−acceptor moieties were discussed and summarized. The discussion was divided into
conclude that the TPA cross section value is linearly proportional four sections by the methods for controlling aromaticity. In the
to the first hyperpolarizability (β).59,60 Actually, these researchers first section, the change of aromaticity by decreasing the
did not directly state the relationship between hyperpolariz- temperature was illustrated. Next, the issue of aromaticity on
abilities or between hyperpolarizability and BLA. However, from protonation and deprotonation was addressed. In the following
their results and logics, the imaginary part of γ, which is the section, the alternation of aromaticity by regulating solvent
second-order hyperpolarizability, is concluded to determine the polarity was discussed. Finally, the reversal of aromaticity in the
TPA cross section. Furthermore, in an earlier study, Brédas et al. excited state of expanded porphyrins was described by analyzing
have reported that BLA is an essential parameter to determine the absorption spectra between the ground and lowest singlet/
the NLO properties shown in conjugated organic systems.61 triplet excited states.
Moreover, experimental results reveal that there is a close
relationship between the aromaticity and TPA cross section
values.62−70 The aromaticity and TPA cross section value are 2. CONTROL OF AROMATICITY BY REGULATING THE
closely related with regard to both ring current and molecular TEMPERATURE
hyperpolarizability. Thus, irrespective of the baseline criterion The aromaticity or antiaromaticity of expanded porphyrins can
chosen for defining the aromaticity, the TPA cross section values be switched by changing the surrounding environment;
can be used as experimental criteria for molecular aromaticity. temperature variation is one of the methods best known to
Consequently, these photophysical features of expanded change the aromaticity of expanded porphyrins. Expanded
porphyrins in the absorption, emission, TA spectra, and TPA porphyrins exhibit a change in aromaticity, in addition to
cross section values, which are dependent on their aromaticity, conformational changes, at low temperature. Experimental
originate from the characteristic electronic structures of results and theoretical calculations completely support the
(anti)aromatic nature. In this context, these photophysical change of aromaticity at low temperature. In this section, the
features can serve as reliable indices for determining the change of aromaticity by decreasing the temperature was
aromaticity, antiaromaticity, and nonaromaticity of expanded discussed.
porphyrins.
2.1. Change of Aromaticity in
The molecular aromaticity of expanded porphyrins can be
meso-Hexakis(pentafluorophenyl) Hexaphyrin at Low
easily controlled by changing the surrounding environment, such Temperatures
as temperature,62−64 viscosity and polarity of solvents,64,65 and
protonation and deprotonation with acids and bases.66−70 In 2008, Kim, Osuka, and co-workers reported temperature-
Furthermore, recently, the aromaticity of expanded porphyrins dependent photophysics of meso-hexakis(pentafluorophenyl)
in the ground state has been reported to be reversed in the [26]- (PF26H) and [28]hexaphyrins(1.1.1.1.1.1) (PF28H)
excited singlet and triplet states.57,71−73 That is, the aromaticity (Chart 1).63 In tetrahydrofuran (THF) at room temperature,
of expanded porphyrins can be easily modulated by changing the the absorption spectrum of PF26H show an intense B-like band
conditions of molecules, indicating that expanded porphyrins are at 571 nm and Q-like bands at 712, 768, 892, and 1018 nm with
an ideal platform for investigating aromaticity switching and low oscillator strength. On the other hand, that of PF28H shows
2260 DOI: 10.1021/acs.chemrev.6b00313
Chem. Rev. 2017, 117, 2257−2312
Chemical Reviews Review

an asymmetric B-like band centered at 596 nm with relatively


broad Q-like bands at 764, 847, 892, and 997 nm (Figure 3).
The absorption spectrum of PF28H at 173 K exhibits a
bathochromic-shifted B-like band at 605 nm accompanied by the

Figure 4. Synchronous (top) and asynchronous (bottom) 2D


absorption correlation spectrum generated from temperature-depend-
Figure 3. Temperature-dependent absorption spectra of PF26H (top)
ent spectral variation of PF28H. Solid and dashed lines represent
and PF28H (bottom) from 293 to 173 K in THF. Reprinted with
positive and negative cross peaks, respectively. Reprinted with
permission from ref 63. Copyright 2009 American Chemical Society.
permission from ref 63. Copyright 2009 American Chemical Society.

growth of a shoulder at 650 nm. PF26H also shows a red-shift in


the B-like band, but the extent of spectral shift is significantly
small as compared to PF28H.
The authors have performed a consistent 2D correlation
spectroscopy to find the origin of distinct spectral evolution in
the absorption spectra of PF28H. In this technique, two
independent variables, providing information on structural
changes according to the temperature change, were used to
plot spectral intensity.74,75 At low temperature, the band at 605
nm in the synchronous spectrum is divided into two distinct
peaks at 593 and 608 nm in the asynchronous spectrum (Figure
4), which supports that at least two structurally different species
of PF28H coexist in the ground state. Especially, the change of
relative band intensity between two separated peaks depending
on temperature represents ground-state equilibrium shift
between the two conformers: one corresponding to the peak at
593 nm is dominant at high temperature, while the other
corresponding to the peak at 608 nm gradually becomes Figure 5. ALS (alternating least-squares)-based SMCR results for
dominant upon lowering the temperature. temperature-dependent absorption spectra of PF28H in THF. Red and
The change of absorption spectra and the corresponding black solid lines indicate the ground-state absorption spectra of the
individual chemical components, and inset shows relative population
population change of individual PF28H conformers with ratio changes from 293 to 173 K. Reprinted with permission from ref 63.
changing temperature were evaluated by a self-modeling curve Copyright 2009 American Chemical Society.
resolution (SMCR) method (Figure 5) in more detail.63 By using
this method, two ground-state absorption spectra, which are
responsible for the two different conformers, were extracted; one band at 589 nm and broad and featureless Q-like bands. As
conformer exhibits a B-like band at 604 nm with Q-like bands at described in the Introduction, expanded porphyrins that have
764 and 890 nm, while the other one exhibits a blue-shifted B-like aromatic character generally exhibit distinct Q-like transitions in
2261 DOI: 10.1021/acs.chemrev.6b00313
Chem. Rev. 2017, 117, 2257−2312
Chemical Reviews Review

the absorption spectra. In contrast, featureless and smeared Q-


like bands are observed in the case of antiaromatic expanded
porphyrins.76,77 In this regard, the two differentiated absorption
spectra revealed by the SMCR analysis finally tell us that Möbius
aromatic and Hückel antiaromatic PF28H conformers, which
have the [4n]π conjugation pathway in common but form
different structures (the former is half-twisted structure while the
latter is planar structure), coexist in equilibrium. Further, by the
temperature-dependent SMCR analysis, population shift be-
tween the two conformers could be quantitatively analyzed. With
decreasing temperature, the relative portion of Möbius aromatic
PF28H increases from 55% to 75%, whereas that of planar
Hückel antiaromatic congener decreases from 45% to 25%,
implying that twisted Möbius aromatic PF28H is energetically
more stable than planar Hückel antiaromatic PF28H in the
ground state.
NMR measurements with varying temperature can give
detailed information on the activation energy barrier associated
with the structural interconversion between planar Hückel and
twisted Möbius PF28H in the ground state. In particular, the
interconversion rate constants (kint) between the two different
conformers can be obtained by analyzing the NMR spectral
changes. At low temperature, structural changes take place more
slowly than in the NMR timescale. Meanwhile, the NMR
spectrum reveals two different sets of signals originating from
two respective nuclei. With increasing temperature, these
processes occur on the NMR timescale, resulting in peak
broadening followed by a change into coalescent spectral shapes. Figure 6. Temperature-dependent absorption and fluorescence spectra
At high temperature where these processes occur more quickly of PF26H (top) and PF28H (bottom) excited by the 442 nm line of a
than the NMR timescale, only a single narrow line is observed at He−Cd laser at 293, 263, 233, 203, and 173 K in THF. Reprinted with
the central position between the two chemical shifts. In reality, permission from ref 63. Copyright 2009 American Chemical Society.
the 13C NMR spectra of PF28H at 173 K exhibit two signal sets,
which are observed at 98 and 106 ppm, but these two peaks distinctly from 358 to 254 cm−1 (∼104 cm−1) (Figure 7). These
gradually broaden and coalesce with increasing temperature. results imply that the conformational relaxations from the
Eventually, at 293 K, only a single peak is observed at 102 ppm. Franck−Condon to relaxed geometry in the S1 state potential
On the basis of these results, the Arrhenius plot of ln kint vs 1/T energy surface are more prominent for PF28H than those for
can be drawn, and the slope of the plot represents the activation PF26H. The relative fluorescence quantum yields (Φf) between
energy barrier of which the value is evaluated to be ∼8 kcal/mol. PF26H and PF28H exhibit interesting features. Assuming that
The same result was also observed by analyzing the 1H NMR the fluorescence quantum yields of PF26H and PF28H at 173 K
spectra. Next, the authors carried out single-point energy are unity, the decreasing ratio of relative Φf for PF28H with
calculation based on the optimized geometry of Hückel temperature is more prominent than that for PF26H [−3.6 ×
antiaromatic and Möbius aromatic PF28H in order to evaluate 10−3 K−1 vs −1.75 × 10−3 K−1], which also supports the
the energy difference between the two conformers. Con- coexistence of various conformers at high temperature in the case
sequently, it was revealed that the total energy of twisted Möbius of PF28H.
aromatic PF28H is lower than that of planar Hückel antiaromatic Femtosecond and nanosecond TA experiments gave fruitful
PF28H by ∼3.7 kcal/mol. information on the excited-state dynamics of PF26H and PF28H
The fluorescence spectra of PF26H and PF28H by varying the depending on their temperature-dependent structural changes.
temperature from 293 to 173 K are shown in Figure 6. Intense (Table 1, Figures 6 and 7).78
peaks are observed at 1032 nm for PF26H and 1034 nm for In the femtosecond TA spectra of PF26H and PF28H, strong
PF28H, with weak vibronic side bands observed around 1250 GSB signals are observed at 565 and 605 nm, corresponding to
nm for both compounds. The full-widths at half-maximum their B-like bands, respectively, with weak ESA bands on both
(FWHMs) of the intense fluorescence bands of PF26H and sides of the GSB bands (Figure 8). The decay profiles of PF26H
PF28H are presumed to be ∼330 and ∼700 cm−1, respectively. at all wavelength regions were well-fitted to a single exponential
With decreasing temperature, the fluorescence spectra com- decay function. Further, their temperature dependences were
monly become sharper, blue-shifted, and enhanced for both nearly negligible, which again indicates that PF26H exists as a
compounds. On the other hand, the degree of change in relative well-defined Hückel aromatic conformer. On the other hand,
fluorescence quantum yields and Stokes shifts is rather different PF28H shows a drastic change in S1-state lifetime and its probe
between PF26H and PF28H. On the basis of the lowest Q-like wavelength as well as temperature dependence. First, with
transitions in absorption, the Stokes shifts of PF26H and PF28H decreasing temperature, the kinetic profiles at 616 nm show a
are evaluated to be around 133 and 358 cm−1 at room single exponential decay, but the decay component considerably
temperature, respectively. The Stokes shifts of PF26H become increases from 186 (293 K) to 260 (173 K) ps. Especially, the
smaller, decreasing from 133 to 108 cm−1 (∼25 cm−1) with kinetic profiles at 638 nm show double-exponential behavior
decreasing temperature, while those of PF28H change more with time constants of ∼17 (30%) and ∼210 (70%) ps at 293 K.
2262 DOI: 10.1021/acs.chemrev.6b00313
Chem. Rev. 2017, 117, 2257−2312
Chemical Reviews Review

Figure 8. Femtosecond TA spectra of PF26H (top) and PF28H


Figure 7. Temperature-dependent changes in Stokes shifts (top) and (bottom) in THF at 173 K. The inset shows the temperature-dependent
relative fluorescence quantum yields (bottom) of PF26H and PF28H. S1-state kinetic profiles from 293 to 173 K in THF. Reprinted with
Reprinted with permission from ref 63. Copyright 2009 American permission from ref 63. Copyright 2009 American Chemical Society.
Chemical Society.

Table 1. Singlet and Triplet (π, π*) Excited-State Lifetimes of of ∼17 and ∼200 ps, is observed at 580 nm. Notably, the
PF26H and PF28H in Temperatures from 293 to 173 K63 contribution of the shorter decay component (44%) when using
PF26H PF28H 580 nm as a pump is greater than that when using 620 nm (22%).
Similar to that observed in the femtosecond TA spectra, strong
temp (K) τS (ps) τT (μs) τS (ps) τS (ps) τT (μs)
GSB and ESA signals are also shown for both PF26H and
293 113 4.8 183 17 (30%), 210 (70%) <0.1 PF28H in the nanosecond TA experiments.82 At 293 K, the
273 121 5.8 179 17 (17%), 246 (83%) 0.15 lifetime of the T1 state of PF26H is determined to be ∼4.8 μs,
253 126 6.5 203 17 (19%), 269 (81%) 0.3 whereas that of PF28H cannot be determined, supposedly due to
233 137 8.7 223 17 (15%), 232 (85%) 1 a short T1-state lifetime within the instrument response time of
213 138 10.5 250 17 (10%), 243 (90%) 7 100 ns. In contrast, at 173 K, the T1-state lifetime of PF28H
193 139 10.9 250 250 30 greatly increases up to 90 μs, which is longer than that of PF26H,
173 140 11.9 260 260 90 11.9 μs (Figure 9). This result is analogous to the trend of
excited-state dynamics observed by femtosecond TA measure-
Interestingly, with decreasing temperature, the contribution to ments and also consistent with the temperature-dependent
the shorter decay component decreases from 30 to 10%, whereas structural interconversion dynamics between Möbius aromatic
that by the longer decay component increases from 70 to 90%. and Hückel antiaromatic PF28H.
The increase in contribution of the longer decay component On the basis of the time-resolved and steady-state
accompanied by the decrease in that of the shorter one with spectroscopic results with varying temperature, the energy
decreasing temperature is consistent with the results obtained by relaxation pathways for PF26H and PF28H at 293 and 173 K can
temperature-dependent SMCR and NMR analyses. This alludes be depicted as shown in Figure 10. The excited-state dynamics in
that the shorter time constant is stemmed from Hückel both S1 and T1 states of PF26H are insensitive to the
antiaromatic conformer but the longer one is responsible for temperature change because of its structurally rigid and planar
Möbius aromatic conformer.79−81 Pump wavelength-dependent [4n+2]π Hückel aromatic nature. On the other hand, PF28H is
femtosecond TA measurements (λpump = 580 and 620 nm) more sensitive to the temperature, resulting from the conforma-
further support this hypothesis. Without reference to the pump tional interconversion processes between the two structures,
wavelength, the kinetic profile at 616 nm exhibits a single [4n]π planar Hückel antiaromatic and [4n]π Möbius aromatic
exponential decay with the time constant of ∼200 ps. On the conformers. Nonplanar porphyrins typically reveal larger Stokes
other hand, biexponential decay profile, with the time constants shifts, broader optical spectra, shorter S1-state lifetimes, and more
2263 DOI: 10.1021/acs.chemrev.6b00313
Chem. Rev. 2017, 117, 2257−2312
Chemical Reviews Review

current, was found to be one of the most suitable parameters


for evaluating the aromaticity of π-conjugated cyclic molecules.52
At the unweighted geometric center of the molecule, Möbius
PF28H exhibits a large negative NICS value (−15.1 ppm). It
shows a large HOMA (0.85) value close to unity and also forms
four frontier orbitals, which is in good agreement with
Gouterman’s four-orbital model of aromatic porphyrins. All
these parameters greatly support the distinct aromatic nature of
Möbius PF28H.
The temperature-dependent TPA cross section values
obtained by femtosecond Z-scan method can give an important
insight into the aromaticity of PF26H and PF28H (Figure
11).87PF26H exhibits significantly higher TPA cross section
values as compared with its congener PF28H at room
temperature. The higher value of TPA cross section of PF26H
relative to PF28H reflects its stronger diatropic ring current.
Regardless of temperature, TPA cross section values of PF26H
remain constant.63 On the other hand, the TPA cross section
value of ∼2100 GM for PF28H at 298 K drastically enhances to
∼9100 GM, which is similar to that for PF26H at 173 K.
Therefore, the temperature-dependent TPA results of PF28H
are in parallel with the temperature-dependent NMR data, again
indicating that the conformational interconversion of PF28H
between planar Hückel antiaromatic and twisted Mö bius
aromatic conformation is totally stopped at 173 K. Considering
the relationship between TPA cross section value and aromaticity
described in the Introduction, the largely enhanced TPA cross
Figure 9. Nanosecond flash photolysis spectra of PF26H (top) and section values of PF28H at low temperature identify that, with
PF28H (bottom) in THF at 173 K. The sample was degassed with argon decreasing temperature, the degree of aromaticity increases due
gases. The inset shows the temperature-dependent T1-state decay to an increase in relative population of Möbius aromatic
profiles from 273 to 173 K. Reprinted with permission from ref 63. conformers. In addition, notably, at 173 K the TPA cross section
Copyright 2009 American Chemical Society.
value of Möbius PF28H is nearly two times higher than those of
group 10 metal-coordinated hexaphyrins.63 The dihedral angle
sensitive temperature-dependent behavior as compared to planar between the most distorted pyrroles is lower in Möbius PF28H
congeners, due to their structural flexibilities. Actually, Holten (∼26.5°) in comparison with that of metal-coordinated
and co-workers have reported that the photophysics of saddled hexaphyrin (39.9−45.8°).11,62 This small dihedral angle permits
and ruffled porphyrins rather differ from those of planar effective and smooth π-conjugation.88 Thus, these results suggest
congeners, such as H2OEP or H2TPP.83−86 Nevertheless, it is
that PF28H formed at low temperature exhibits less-strained
important to note that the lifetimes of twisted Möbius aromatic
Möbius topology as compared to that of metal-coordinated ones.
PF28H, despite its nonplanar geometry, in the S1- and T1-excited
states are quite longer than those of planar PF26H having Hückel Moreover, the TPA cross section values of Möbius aromatic
aromaticity. PF28H are similar to those of PF26H. This also indicates that
For explaining these interesting observations for PF28H, two the degree of aromaticity of Möbius aromatic PF28H is
viewpoints were proposed: one is the increase of conformational comparable to that of Hückel aromatic PF26H.
rigidity, and the other is the degree of aromaticity. Because the To design stable Möbius aromatic systems, an appropriate size
excited-state lifetimes of PF28H increase with decreasing of π-conjugated macrocycles is a crucial factor that must be taken
temperature, the potential energy surface of Möbius aromatic into account. Herges and co-workers have reported that from
PF28H is suggested to possess a curvature narrower than that of small twisted cyclic systems it is hard to synthesize a stable
Hückel aromatic PF26H in both ground and S1 states. That is, Möbius structure because of their large structural constraint.89
they exhibit a narrower potential energy surface having larger On the other hand, in large [4n]annulene, ring strain is greatly
force constants, which is the characteristic of Hessian. Thus, the reduced. However, due to its very flexible structure, it would
Mö bius structure of PF28H experiences more restricted naturally return to the Hückel topology, which has small strain.
molecular motion. In other words, with decreasing temperature,
Thus, it is difficult to accomplish the locking of a half-twisted
PF28H becomes locked into the rigid and stable Möbius
topology. According to previously reported studies on diverse Möbius structure. On the basis of various experiments and
expanded porphyrins, most of the aromatic expanded porphyrins quantum calculations, PF28H can be believed to have a smooth
exhibit longer S1- and T1-state lifetimes than their antiaromatic structure and a proper ring size for realizing a stable Möbius
congeners. Therefore, the degree of aromaticity is believed to be aromatic system with optimum π-conjugation. Therefore, with
an important factor that is strongly related with the excited-state decreasing temperature, a delicate balance between the ring
dynamics in expanded porphyrins. The NICS value, which strain induced by the twisted Möbius topology and the energetic
reflects magnetic (de)shielding effects of π-electronic ring stabilization by π-conjugation occurs in Möbius PF28H.
2264 DOI: 10.1021/acs.chemrev.6b00313
Chem. Rev. 2017, 117, 2257−2312
Chemical Reviews Review

Figure 10. Schematic diagram for the energy-relaxation pathways of PF26H and PF28H. Activation energy is calculated from temperature-dependent
13
C and 1H NMR spectra. Reprinted with permission from ref 63. Copyright 2009 American Chemical Society.

Figure 11. Open aperture Z-scan trace for PF26H (left) and PF28H (right) in THF under variable temperature with excitation at 1200 nm. The sample
concentrations for PF26H and PF28H are 0.3 and 0.5 mM, respectively. The red lines are the best fitted curves of experimental data. Reprinted with
permission from ref 63. Copyright 2009 American Chemical Society.

2.2. Change of Aromaticity in TFM26H and TFM28H are believed to be Hückel aromatic
meso-Hexakis(trifluoromethyl) Hexaphyrins at Low and antiaromatic species, respectively.61,90 Nevertheless, both
Temperatures experimental and theoretical studies were hard to match this
Similar temperature-dependent behaviors with regard to the simple assignment.88,91,92
photophysical properties of meso-hexakis(trifluoromethyl) [26]- Figure 12 shows the absorption and emssion spectra of
and [28]hexaphyrins, TFM26H and TFM28H, respectively, TFM26H and TFM28H measured in toluene and dichloro-
were also reported by the groups of Kim and Osuka in 2011 methane at 298 K. The spectral characteristics of TFM26H and
(Chart 2).90 They are intriguing because both hexaphyrins are TFM28H are quite analogous to those observed for PF26H and
figure-eight geometry, despite their different numbers of π PF28H.60,93 The absorption spectrum of TFM26H in toluene
electrons. As both compounds exhibit double half-twists, shows sharp and intense B-like band with four Q-like bands.
2265 DOI: 10.1021/acs.chemrev.6b00313
Chem. Rev. 2017, 117, 2257−2312
Chemical Reviews Review

Chart 2. Molecular Structures of TFM26H and TFM28H Based on X-ray Crystallographic Data (Reprinted with Permission from
Ref 64; Copyright 2011 American Chemical Society)

Figure 13. Absorption spectra of TFM26H and TFM28H in CH2Cl2 as


plotted by their extinction coefficients. Reprinted with permission from
ref 64. Copyright 2011 American Chemical Society.

from 113 to 293 K display different spectral features. For


TFM26H, with decreasing temperature, the intensities of all
bands increase and band narrowing occurs in the B-like bands,
which is probably due to the enhanced structural rigidity (Figure
14). On the other hand, as shown in Figure 15, significantly
Figure 12. (a) Absorption spectra of TFM26H (black line) and different behavior is observed in the temperature-dependent
TFM28H (red line) in toluene (solid line) and CH2Cl2 (dashed line). absorption spectra of TFM28H. With decreasing temperature,
(b) NIR fluorescence spectra of TFM26H (black dashed line) and the intensities of all bands gradually increase, which is similar to
TFM28H (red dashed line) with photoexcitation at 532 nm in toluene. that observed for TFM26H. Exceptionally, the intensity of the
Absorption and emission spectra are normalized to the highest band around 500 nm considerably reduces. On the basis of the
absorption and lowest absorption bands, respectively. The absorption
change of absorption spectra of TFM28H with varying solvent
bands in the 700−1200 nm region are magnified for clarity. Reprinted
with permission from ref 64. Copyright 2011 American Chemical polarity,61 it can be considered that the enhanced and reduced
Society. absorption bands with decreasing temperature originate from the
Hückel antiaromatic and Möbius aromatic conformers, respec-
tively. In other words, the conformational equilibrium dynamics
Furthermore, TFM26H emits strong fluorescence in the NIR controlled by the temperature also occur in the case of TFM28H.
region, which is characteristic of typical aromatic porphyr- The Möbius aromatic and Hückel antiaromatic conformers are
inoids,38,94 indicating that it has strong aromaticity. The ground dominant at low and high temperatures, respectively, again
state optical properties of TFM28H in toluene also exhibit verifying that the former is energetically more stable than the
similar features to those observed for PF28H,93 such as an latter. This observation follows the same trend of temperature-
intense B-like band with four Q-like bands in absorption and also dependent behavior of PF28H,59,60 which we discussed in the
a strong fluorescence in the NIR region. The only difference previous section.
between these two compounds is a weak absorption peak at The femtosecond TA measurements of TFM26H and
∼500 nm of TFM28H. As compared with those of vinylene- TFM28H greatly support the temperature-dependent, steady-
bridged [26]- and [28]hexaphyrins,54 the broader absorption state measurements. In toluene, the excited-state dynamics of
spectra and reduced extinction coefficients of both TFM26H and TFM26H exhibit a trend similar to those observed for typical
TFM28H suggest that the figure-eight geometry has large aromatic expanded porphyrinoids (Figure 16).38 Strong negative
flexibility (Figure 13). bands of GSB signals and broad and positive ESA signals are
Despite the similar absorption spectra of TFM26H and observed, where the whole shapes of spectra do not differ within
TFM28H, their absorption spectra in the temperature range 1 ns window. Two lifetime components of 1 and 55 ps are
2266 DOI: 10.1021/acs.chemrev.6b00313
Chem. Rev. 2017, 117, 2257−2312
Chemical Reviews Review

Figure 14. (a) Temperature-dependent absorption spectra and (b)


logarithmic plot of absorbance change as a function of temperature at
specific wavelengths of TFM26H in 2-methyltetrahydrofuran (2- Figure 15. (a) Temperature-dependent absorption spectra and (b)
MTHF). In (a), arrows indicate the direction of decreasing temperature. logarithmic plot of absorbance change as a function of temperature at
Reprinted with permission from ref 64. Copyright 2011 American specific wavelengths of TFM28H in 2-MTHF. In (a), arrows indicate
Chemical Society. the direction of decreasing temperature. Reprinted with permission
from ref 64. Copyright 2011 American Chemical Society.

observed in both GSB (570 nm) and ESA (600 nm) bands. The
former time constant is attributed to the internal conversion time constants, 1 and 10 ps, are quite similar in comparison to
from the B-like state to the lowest excited state and/or vibrational those observed for vinylene-bridged [28]hexaphyrins (0.53 and
cooling processes. The latter one with large contribution can be 8.6 ps in toluene), which is the Hückel antiaromatic compound.54
attributed to the S1-state lifetime. This result indicates that the TA spectra after photoexcitation at
Compared with PF26H and vinylene-bridged [26]- 500 nm are mainly governed by the excited state of the Hückel
hexaphyrins (∼100 and ∼280 ps, respectively),54,60 the S1-state antiaromatic conformer.
lifetime of TFM26H is slightly shorter, which is presumably due In contrast, the longer transient species having lifetime
to the structural flexibility of the figure-eight structure, as constants of 255 ps and >5 ns are responsible for the S1- and
discussed earlier. For TFM28H, pump-wavelength-dependent T1-state lifetimes of the Möbius conformer of TFM28H,
TA measurements were carried out for the selective excitation of respectively, because both time constants and spectral shapes
Hückel and Möbius conformers (Figure 17). Notably, accurate are analogous to those of the Möbius aromatic hexaphyrins,
selective excitation is impossible because, in the B-like band PF28H.60 When the pump wavelength was changed to 560 nm
region, there is a considerable spectral overlap between both corresponding to the B-like band for the Möbius conformer of
conformers.61 TFM28H, some different excited-state dynamics were observed
Nonetheless, spectral changes are distinctly observed in the (Figure 17b, d, f). Not only the decay time constants but also the
pump-wavelength-dependent TA spectra. In the TA spectra after spectral shapes of DAS are similar, while the magnitude of the
photoexcitation at 500 nm, two strong GSB bands are observed GSB bands for the Hückel and Möbius conformers, particularly
at 500 and 560 nm (Figure 17a). Additional GSB and ESA bands their DAS of ∼1 and 250 ps, are dissimilar (Figure 17c, d). These
are also observed at 507 and 650 nm but quickly decay within a differences distinctly indicate that the relative population of the
few picoseconds. This is almost 20 times shorter than the decay excited Hückel conformer is greater than that of the Möbius
time constant of the GSB band observed at 565 nm (Figure 17e). conformer, which can be regulated by selective excitation. These
Four transient species corresponding to 1, 10, 255, and >5000 ps changes in the excited population controlled by the excitation
are extracted in the decay-associated spectra (DAS) of TFM28H wavelength again identify the coexistence of two different
obtained by global analysis (Figure 17c). Notably, temporal conformers in TFM28H depending on their aromaticity.
evolution and spectral shapes of the TA spectra for the faster two Because the optically dark state of the antiaromatic conformer
2267 DOI: 10.1021/acs.chemrev.6b00313
Chem. Rev. 2017, 117, 2257−2312
Chemical Reviews Review

Figure 16. TA spectra (top) of TFM26H in toluene (a, b) and CH2Cl2 (c, d) with the photoexcitation at 580 nm and temporal profiles (bottom) probed
at 570 (black circles) and 660 nm (red circles) retrieved from (a) and (c), respectively, with least-squared fitting curves. Reprinted with permission from
ref 64. Copyright 2011 American Chemical Society.

Figure 17. TA spectra (left), decay-associated spectra (middle), and temporal profiles (right) of TFM28H in toluene with photoexcitation at 500 (a, c,
and e) and 560 nm (b, d, and f). Reprinted with permission from ref 64. Copyright 2011 American Chemical Society.

can serve as a ladder state during the relaxation processes, the


time constant of 10 ps for the Hückel antiaromatic conformer in the ground state, which is governed by the energy-gap law
TFM28H stems from the faster internal conversion process to (Figure 18).95
2268 DOI: 10.1021/acs.chemrev.6b00313
Chem. Rev. 2017, 117, 2257−2312
Chemical Reviews Review

Figure 18. Schematic diagram for energy relaxation dynamics of (a) TFM26H and (b) TFM28H in toluene. The values in parentheses were determined
in CH2Cl2 solution. VR and IC represent the vibrational-relaxation and internal-conversion processes, respectively. Reprinted with permission from ref
64. Copyright 2011 American Chemical Society.

As mentioned above, 1H NMR spectroscopy is one of the mixture experiencing conformational interconversion splits into
optimum methods for investigating molecular aromaticity.53,96 two structurally different conformers. At high temperature, >300
Temperature-dependent 1H NMR spectra can give a deeper K, a strong paratropic ring current is observed in the simulated
1
insight into the structural information and equilibrium dynamics H NMR spectrum, verifying that the Hückel conformer is
of the two structures in accordance with their aromaticity (Figure antiaromatic. On the other hand, at a temperature below 150 K,
19). the diatropic ring current becomes prominent, which indicates
that the Möbius conformer is aromatic. In the case of TFM26H,
temperature dependence is not observed. On the basis of the
temperature-dependent 1H NMR spectra of TFM28H, a
thermodynamic model can be utilized by using the van’t Hoff
equation.27 Suppose the following conformational equilibrium
between two different compounds:
H↔M
Here, species H and M designate the Hückel and Möbius
conformers, respectively, in TFM28H. For a faster conforma-
tional interconversion process than the 1H NMR timescale, the
observed chemical shift is an average chemical shift of the two
isomers during the NMR timescale.97,98
For these reasons, the observed chemical shift δi(T) of proton i
at a given temperature T can be described by the following
equation:
δ i(T ) = x H(T )δ Hi + xM(T )δMi (1)
1
Figure 19. Temperature-dependent H NMR spectra of TFM28H in Here, xH(T) and xM(T) represent the mole fractions of the
THF-d8. Reprinted with permission from ref 64. Copyright 2011 Hückel and Möbius conformers at temperature T, and δiH and
American Chemical Society. δiM represent the chemical shifts of the ith protons for the Hückel
and Möbius conformers, respectively. Moreover, on the basis of
With decreasing temperature from 298 to 173 K, dramatic the van’t Hoff equation, the equilibrium constant K(T) for this
changes are observed in the 1H NMR spectra of TFM28H as reversible process can be estimated by
follows: (i) the most deshielded signal (14.43 ppm) of the NH
proton is significantly upfield-shifted (to 7.08 ppm) and also the
[M]eq x M(T ) ⎛ ΔS ΔH ⎞⎟
K (T ) = = = exp⎜ −
intersection points between two NH protons emerge between [H]eq x H(T ) ⎝ R RT ⎠ (2)
213 and 233 K; (ii) the complicated signal, which is originated
from the six β-CH protons (7.22 ppm), is divided into four Using eq 2 and xH(T) + xM(T) = 1 under any condition, eq 1 can
deshielded peaks (from 7.98 to 8.33 ppm) and two shielded be rewritten as
peaks (4.11 and 4.42 ppm); and (iii) the signal intensities of all K (T )
peaks increase. These spectral evolutions can be explained with δ i(T ) = (δMi − δ Hi ) + δ Hi
regard to repressed structural dynamics at low temperature; a 1 + K (T ) (3)

2269 DOI: 10.1021/acs.chemrev.6b00313


Chem. Rev. 2017, 117, 2257−2312
Chemical Reviews Review

Finally, ΔH, ΔS, δiH, and δiM can be determined by unweighted isomers is rather sensitive to the polarity of the solvent, the
least-squares fitting by minimizing quantity R as follows:99 temperature-dependent 1H NMR spectra of TFM28H were also
performed in toluene-d8. The overall change in the 1H NMR
R= ∑ (δ i(Tj)calc + δ i(Tj)expt )2 spectra of TFM28H in toluene-d8 resembles that of the 1H NMR
i,j (4) spectra in THF-d8. However, slightly smaller ΔS (−35.5 J·K−1·
Here, i and j are the index peaks and temperature values, mol−1) and ΔH (−9.9 kJ·mol−1) were observed.61 Interestingly,
respectively. The results of nonlinear curve-fitting analysis with the two NH proton peaks were not intersected in toluene-d8
eqs 3 and 4 by using the temperature-dependent 1H NMR from 183 to 298 K. The extrapolated chemical shifts for the
spectra of TFM28H in THF-d8 are illustrated in Figure 20, and Hückel and Möbius conformers could also be estimated based on
several parameters are displayed in Table 2. the temperature-dependent NMR spectra of TFM28H. Because
the number of single peaks used in the analyses of all
temperature-dependent NMR data is half of the total number
of protons in hexaphyrin, the Hückel and Möbius conformers
should have C2 symmetry (Table 2). This requirement is fulfilled
in the case of Hückel conformer, while the Möbius conformer has
a low symmetry affording complicated calculations for the 1H
NMR spectra. However, this is different from the structure
predicted from temperature-dependent NMR data (Table 3 and
Figure 21).61 This discrepancy implies that even at 173 K
dynamic interconversion processes occur as rapidly as ever.
An analogous dynamic equilibrium between the Hückel
antiaromatic and Möbius aromatic conformers was also observed
in di-p-benzi-[28]hexaphyrins reported by Latos-Grażyński et
al.23 In their case, it is important to note that the Möbius aromatic
conformer exhibits C2 symmetry based on both quantum
mechanical calculation results and X-ray crystallographic data.
Figure 20. Experimental (closed circles) and fitted (solid lines) 1H Surprisingly, the spectral evolution in the 1H NMR spectra of
NMR chemical shifts of TFM28H in THF-d8 as a function of TFM28H with varying temperature is analogous to those of
temperature. Reprinted with permission from ref 64. Copyright 2011 PF28H in 1,1,2,2-tetrachloroethane-d2 (198−413 K).59 In the
American Chemical Society. case of PF28H, the crossing temperature (Tcross) is estimated to
be ∼400 K, affording an estimated critical temperature (Tunity) of
The determined thermodynamic parameters of ΔS (−57.2 J· ∼380 K, which is significantly greater than that of TFM28H.
K−1·mol−1) and ΔH (−12.4 kJ·mol−1) suggest that the proposed Notably, the 1H NMR peaks are divided and become more
mechanism is not only exothermic but also spontaneous. That is, complex with the loss of C2 symmetry at a temperature below 200
the Möbius conformation is more stable than the Hückel K. Nevertheless, at room temperature, it can be said that the
conformation. Furthermore, as the equilibrium between the two Möbius aromatic conformer is dominant at equilibrium and rapid

Table 2. Experimental and Simulated 1H NMR Band Positions of TFM28H in THF-d8 and Toluene-d864

chemical shifts
solvent temp (K) 2(NH) 2(NH) 2(β-CH) 2(β-CH) 2(β-CH) 2(β-CH) 2(β-CH) 2(β-CH) K (T)
THF-d8 298 14.43 (14.54) 12.42 (12.34) 7.22 (7.33) 7.22 (7.17) 7.22 (7.32) 7.22 (7.18) 7.22 (6.96) 7.22 (7.2) 0.15
273 13.91 (13.9) 12.23 (12.22) 7.42 (7.42) 7.25 (7.26) 7.42 (7.38) 7.25 (7.25) 6.63 (6.73) 6.92 (6.94) 0.24
253 13.17 (13.11) 12.05 (12.08) 7.6 (7.53) 7.36 (7.39) 7.49 (7.45) 7.3 (7.33) 6.34 (6.46) 6.6 (6.61) 0.37
233 11.97 (12.02) 11.82 (11.88) 7.75 (7.69) 7.54 (7.56) 7.6 (7.56) 7.43 (7.45) 5.97 (6.05) 6.14 (6.13) 0.61
213 10.47 (10.44) 11.55 (11.6) 7.93 (7.9) 7.78 (7.8) 7.74 (7.7) 7.61 (7.61) 5.48 (5.52) 5.51 (5.5) 1.12
193 8.66 (8.68) 11.27 (11.29) 8.14 (8.15) 8.07 (8.07) 7.78 (7.86) 7.8 (7.8) 4.9 (4.9) 4.76 (4.77) 2.31
173 7.08 (7.11) 11.08 (11.01) 8.33 (8.37) 8.33 (8.31) 8.02 (8.01) 7.98 (7.97) 4.42 (4.34) 4.11 (4.11) 5.64
Hückel 15.90 12.58 7.14 6.95 7.19 7.04 7.44 7.77
conformer
Möbius 5.55 10.73 8.59 8.55 8.15 8.13 3.79 3.46
conformer
toluene- 298 11.3 (11.32) 11.63 (11.62) 7.29 (7.31) 7.15 (7.16) 6.91 (6.92) 6.91 (6.94) 5.51 (5.48) 5.24 (5.22) 0.76
d8 273 10.27 (10.28) 11.36 (11.36) 7.37 (7.36) 7.26 (7.26) 7 (6.99) 6.94 (6.94) 5.17 (5.16) 4.83 (4.83) 1.09
253 9.32 (9.31) 11.12 (11.12) 7.42 (7.41) 7.36 (7.35) 7.06 (7.06) 6.97 (6.97) 4.85 (4.87) 4.46 (4.47) 1.54
233 8.29 (8.27) 10.86 (10.86) 7.47 (7.47) 7.48 (7.45) 7.14 (7.13) 7 (6.99) 4.52 (4.55) 4.07 (4.09) 2.31
213 7.27 (7.23) 10.6 (10.6) 7.53 (7.52) 7.56 (7.55) 7.21 (7.21) 7.03 (7.02) 4.19 (4.24) 3.69 (3.7) 3.74
193 6.29 (6.31) 10.37 (10.37) 7.57 (7.57) 7.63 (7.64) 7.27 (7.27) 7.04 (7.05) 3.88 (3.96) 3.34 (3.36) 6.67
183 5.89 (5.92) 10.28 (10.27) 7.58 (7.59) 7.67 (7.68) 7.29 (7.3) 7.05 (7.06) 3.97 (3.84) 3.25 (3.21) 9.34
Hückel 16.25 12.86 7.04 6.69 6.57 6.79 6.97 7.05
conformer
Möbius 4.81 10.00 7.65 7.78 7.38 7.08 3.50 2.80
conformer

2270 DOI: 10.1021/acs.chemrev.6b00313


Chem. Rev. 2017, 117, 2257−2312
Chemical Reviews Review

Table 3. Measured and Calculated Chemical Shifts in 1H NMR Spectra of Two Conformers in TFM28H64
δ (Hückel conformer) [ppm] δ (Möbius conformer) [ppm]
signal THF-d8 toluene-d8 GIAOa THF-d8 toluene-d8 GIAOa
2(NH) 15.90 16.25 19.61, 19.64 5.52 4.81 −0.14, 0.17 (inner 2(NH))
2(NH) 12.58 12.86 15.34, 15.36 10.73 10.00 3.13(NH), 13.34 (NH)
2(β-CH) 7.14 7.04 7.08, 7.04 8.59 7.65 7.90−9.50 (outer 8(β-CH))
2(β-CH) 6.95 6.69 5.84, 5.90 8.55 7.78
2(β-CH) 7.19 6.57 6.95, 6.97 8.15 7.38
2(β-CH) 7.04 6.79 6.38, 6.43 8.13 7.08
2(β-CH) 7.44 6.97 8.77, 8.77 3.79 3.50 4.51, 5.14 (2(β-CH))
2(β-CH) 7.77 7.05 8.30, 8.35 3.46 2.80 −3.68, −4.15 (2(β-CH))
a
GIAO = gauge-including atomic orbital calculation.

omatic and aromatic species can be influenced by the movement


of a p-benzene component.53 Even though a detailed mechanism
of the interconversion processes could not be decided yet, it is
believed that a large conformational change, triggered by bond
rotation occurring through two Cα−Cmeso bonds, is necessary for
the structural change between the Hückel antiaromatic and
Möbius aromatic conformers (Figure 22).

Figure 21. Optimized structures and molecular orbital densities of the


lowest π-MOs for Hückel (left side with 161st MO) and Möbius
conformers (right side with 159th and 162nd MOs) at B3LYP/6-
31G(d) level. Reprinted with permission from ref 64. Copyright 2011
American Chemical Society.

interconversion results in the C2 symmetry in the 1H NMR


spectra. Figure 22. Schematic picture for structural change from Hückel
This observation indicates that the intrinsic 1H NMR spectra conformer to Möbius conformer of TFM28H. Blue- and red-colored
of the Möbius aromatic conformer can be observed at a pyrrole moieties represent different sides of the ring. Orange-colored
temperature significantly lower than 173 K where the solvent is pyrroles are vertically arranged with respect to the paper plane. It is
noted that two green arrows meaning main rotational axes for
completely frozen. According to all temperature-dependent
isomerization have the same direction of rotation. Reprinted with
NMR data of three types of [28]hexaphyrins, their conforma- permission from ref 64. Copyright 2011 American Chemical Society.
tional equilibrium enormously relies on external factors, e.g.,
solvent and temperature. The Tunity value increases in ascending
order of TFM28H (216 K in THF-d8), di-p-benzi[28]- In summary, Hückel aromatic TFM26H exists as a figure-eight
hexaphyrins (330 K in CDCl3), and PF28H (380 K in 1,1,2,2- conformer in solution, while TFM28H exists as the two
tetrachloroethane-d2), which identifies that the activation energy structural isomers under dynamic equilibrium, and their
barrier for the conformational interconversion processes equilibrium is rather sensitive to temperature. For TFM28H,
between the Möbius aromatic and Hückel antiaromatic con- one isomer showing a short S1-state lifetime around 10 ps is
formers increases in the order stated above.53 The authors believed to correspond to the Hückel antiaromatic species having
suggested that a nearly barrierless activation energy for the a flexible figure-eight structure, while the other conformer is
structural changes of TFM28H originates from the fact that the believed to correspond to the Möbius aromatic hexaphyrin with
figure-eight structure of TFM28H with stretically less-hindered twisted topology, exhibiting a long S1-state lifetime (250 ps).
trifluoromethyl substituents is structurally more flexible as From these results, it was delineated that the control of
compared to di-p-benzi[28]hexaphyrins and PF28H.53 Similarly, conformational interconversion between aromatic and antiar-
the activation energy barrier declines with increasing solvent omatic mixtures by regulating temperature can afford significant
polarity in TFM28H on the basis of its solvent-dependent 1H insight into the relationship between the molecular structure and
NMR and absorption spectra. For di-p-benzi[28]hexaphyrins, the chemical properties in aromatic and antiaromatic expanded
the conformational interconversion process between antiar- porphyrinoids.
2271 DOI: 10.1021/acs.chemrev.6b00313
Chem. Rev. 2017, 117, 2257−2312
Chemical Reviews Review

3. CONTROL OF AROMATICITY BY PROTONATION two expanded porphyrins possess 30 and 38 π electrons


AND DEPROTONATION satisfying the Hückel [4n+2] counting rule, they have distorted
The NH protons in expanded porphyrins significantly affect the molecular structures. In this regard, a comparative analysis
determination of their molecular structures. By modifying the between PF30h and PF38N in neutral and protonated forms
number of protons of expanded porphyrins by protonation or (PF30hp and PF38Np) leads to in-depth comprehension for the
deprotonation, molecular structures would change from a photophysical properties and aromaticity triggered by proto-
twisted form to an unfolded form, which plays a key role in the nation-induced structural changes.67
formation of π-conjugation pathway and change of aromaticity. Although PF30h and PF38N in their neutral forms possess
In this regard, the protonation and deprotonation of expanded [4n+2]π electrons in the core cyclic π-delocalization for
porphyrins are considered to be easy and efficient for modifying satisfying the Hückel [4n+2] rule, their 1H NMR spectra indicate
the aromaticity of expanded porphyrins. weak aromatic character,100,101 which indicates that their
distorted figure-eight conformations give rise to a weak aromatic
3.1. Control of Aromaticity by Protonation
nature. For freebase PF38N, its X-ray structure clearly supports
3.1.1. Effects of Protonation in [4n+2]π Expanded the weak aromaticity arising from distorted sturctures. Here, a
Porphyrins. In this section, the effect of protonation on notable behavior of PF30h is a spectral change in the 1H NMR
photophysical properties of meso-aryl-substituted expanded spectrum upon the addition of trifluoroacetic acid (TFA). As
porphyrins are described with two expanded porphyrins, meso- compared to the 1H NMR spectrum of its neutral form,
hexakis(pentafluorophenyl) [30]heptaphyrin (1.1.1.1.1.1.0) and protonated PF30h (PF30hp) displays a strong diatropic ring
meso-hexakis(pentafluorophenyl) [38]nonaphyrin current, reflecting the strong aromatic character. Because the X-
(1.1.0.1.1.0.1.1.0) (PF30h and PF38N, respectively; Chart ray crystallography of PF30hp indicates a planar structure with
3),100,101 focusing on revealing the relationship between the three inverted pyrrole rings, the intensified aromatic character
seems to arise from a conformational change from the twisted to
Chart 3. Molecular Structures of PF30h and PF38N planar conformations.100 The 1H NMR spectrum of protonated
(Reprinted with Permission from Ref 67; Copyright 2009 PF38N (PF38Np) shows a large diatropic ring current, which is
American Chemical Society) also attributed to a conformational change upon the addition of
methanesulfonic acid (MSA) (Figure 23). On the other hand, the
β-CH proton peaks are observed at 4−8 ppm in the 1H NMR
spectrum of PF38N. With the addition of MSA, the 1H NMR
spectrum of PF38Np displays a series of peaks in the range of −7
to 18 ppm.101
Furthermore, the 1H−1H-COSY spectroscopic results of
PF38Np indicate that the counteranion of MSA is restricted in
the cavity of the core nonaphyrin macrocycle under the
paratropic ring-current effect of PF38Np.67 Although it was
photophysical properties and the structural planarity in not possible to further confirm its structure by X-ray
connection with their aromaticity. These two porphyrins contain crystallography analysis, these NMR data provide distinct
the identical meso-peripheral substituents and are composed of conformational information that PF38Np is more planar and
seven and nine pyrroles, respectively. Moreover, even though the has intensified aromatic character compared to its neutral

Figure 23. 1H NMR spectrum of PF38N with 2 equiv of MSA in THF-d8: downfield region (top) and upfield region (down). Reprinted with permission
from ref 67. Copyright 2009 American Chemical Society.

2272 DOI: 10.1021/acs.chemrev.6b00313


Chem. Rev. 2017, 117, 2257−2312
Chemical Reviews Review

Figure 24. Spectrophotometric titration of PF30h (a, b) with TFA in CHCl3 and (c, d) with MSA in CH2Cl2. Reprinted with permission from ref 67.
Copyright 2009 American Chemical Society.

congener PF38N. On the basis of these conformational changes


by the 1H NMR spectroscopic data, the twisted and relatively
planar structures for PF30h and protonated PF38Np are
suggested.
The stepwise spectral changes in their absorption spectra by
the addition of acid also delineate the protonation effect in detail.
The acid titration absorption spectra of PF30h with TFA exhibit
two-step spectral changes with distinct isosbestic points (Figure
24). Upon increasing TFA concentration ranging from 0 to 2.1 ×
10−2 M, the B- and Q-like bands are red-shifted with an
appearance of additional longer wavelength bands (Figure 24a).
With increasing TFA concentration up to 2.3 × 10−1 M, the B-
like band is more intensified and blue-shifted (Figure 24b). Even
though the Q-like bands display entangled spectral changes, the
four absorption bands in the NIR region become well-defined
and intense. This change of absorption spectral features during
titration with TFA illustrates the conformational changes of
PF30h triggered by protonation. Also, the spectrophotometric
titration of PF30h with MSA in CH2Cl2 shows similar spectral
changes with two-stepwise processes (Figure 24c, d). Thus, the
addition of MSA to the CH2Cl2 solution of PF30h results in the
same conformational changes as those observed by the addition
of TFA, where the only difference is that the required amount of
MSA for protonation (6.6 × 10−5 M) is significantly less than that
of TFA, representing the higher acidity and stronger binding Figure 25. Spectrophotometric titration of PF38N with MSA in
ability of MSA. The acid titration for PF38N with the addition of CH2Cl2. Reprinted with permission from ref 67. Copyright 2009
MSA also displays the absorption spectral changes. In American Chemical Society.
spectrophotometric titration, intensified B- and Q-like bands
with simultaneous small red-shift are observed in the absorption These neutral and protonated expanded porphyrins display
spectrum of PF38N (Figure 25). Notably, compared to PF30h, significant different features in their absorption spectra (Figures
the spectral change of PF38N occurs at the addition of a large 24 and 26). Compared to the absorption spectrum of PF38N,
amount of MSA (2.9 × 10−1 M). This demonstrates that the that of PF38Np shows no significant absorption bands in the
highly twisted conformation of PF38N restricts appending range of 300−500 nm. The similar spectral change is also
protons to pyrrolic nitrogen atoms, owing to steric hindrance and observed in the absorption spectrum of PF30hp (by the addition
more effective intramolecular hydrogen bonds, thereby decreas- of TFA). In addition, by protonation, PF30h and PF38N show
ing the proton affinity of imine nitrogen atoms. narrower and enhanced B-like bands with negligible spectral
2273 DOI: 10.1021/acs.chemrev.6b00313
Chem. Rev. 2017, 117, 2257−2312
Chemical Reviews Review

(Figure 27). The reduced Stokes shifts by protonation illustrate


that reorganization energy under solvent condition changes by

Figure 26. Steady-state absorption (black) and fluorescence (red) Figure 27. Steady-state absorption (black) and fluorescence spectra
spectra of PF30h, PF38N, and their protonated forms (PF30hp and (red) of neutral PF30h (top) and its protonated form PF30hp
PF38Np, respectively) upon addition of MSA in CH2Cl2. Reprinted (bottom) with TFA in CHCl3. Reprinted with permission from ref 67.
with permission from ref 67. Copyright 2009 American Chemical Copyright 2009 American Chemical Society.
Society.

protonation. According to the reported study for the protonation


shifts (Figure 26). By the addition of acid, the molar of porphyrins,86 this change in Stokes shift indicates the
absorptivities of PF30h and PF38N increase more than three protonation-induced conformational changes of PF30h and
times. Simultaneously, their FWHMs of the B-like bands PF38N, especially from twisted to planar conformation.
decrease. Because the protonation triggers the change of conformation,
Because expanded porphyrins having symmetrical structures electronic structure, and aromaticity, PF30h and PF38N show
generally display narrow absorption bands, these absorption different excited-state dynamics upon the addition of acid. A set
spectral features reflect that protonation induces conformational of TA spectra for both neutral and protonated speceis well-
changes of PF30h and PF38N from twisted to substantially describe such change in the excited-state dynamics (Figure 28).
planar structures. Notably, PF30hp and PF38Np exhibit strong In femtosecond TA measurement, PF30h and PF30hp with
Q-like bands, while weak and ill-defined Q-like bands are MSA show single exponential decay, and the S1-state lifetimes are
observed for PF30h and PF38N. As described in the estimated to be ∼47 and 400 ps (average values from the GSB
Introduction, in-depth analyses for the relationship between and ESA signals), respectively (Table 5). Similarly, PF30hp with
the photophysical properties and aromaticity of expanded TFA exhibits a singlet excited-state lifetime but significantly
porphyrins allow that their optical spectral features provide elongated compared to that of its neutral form (∼167 ps). A
indices for determining their aromatic nature. On the basis of similar change of singlet excited-state lifetime by protonation is
these absorption spectral features of expanded porphyrins, weak obtained in the TA decay profiles of PF38N (Figure 29), where
aromatic character of PF30h and PF38N is assigned by their the singlet excited-state lifetimes are evaluated as ∼18 ps for
broad, smeared, and weak Q-like bands, which is in accordance PF38N and ∼44 ps for PF38Np (average values of GSB and ESA
with their 1H NMR spectroscopic results, even though they have signals).
π electrons meeting the [4n+2] rule for Hückel aromaticity. In In addition, their triplet excited-state lifetimes by nanosecond
other words, reduced aromaticity is attributed to the distorted TA spectroscopy exhibit similar changes to their S1-state
conformations of PF30h and PF38N. Their fluorescence spectra lifetimes. Upon protonation with MSA, the triplet excited-state
in the NIR region also display spectral changes upon acid lifetimes of 3.1 and 0.2 μs for PF30h and PF38N are increased to
titration (Figure 26). PF30h and PF38N show a blue-shift of 6.1 and 6.2 μs, respectively (Tables 4 and 6). PF30hp by TFA
fluorescence emission bands, which is well-matched with their addition shows 2.0 μs as a triplet excited-state lifetime, whereas
blue-shift of lowest-energy Q-like bands. Even though the neither GSB nor ESA signal is observed for PF30h in CHCl3,
magnitude of blue-shifts for the lowest-energy Q-like absorption probably due to less-efficient intersystem crossing and rapid
bands of PF30h and PF38N is not the same as that of their triplet state decay of less than a few tens of nanoseconds (Table
fluorescence, they exhibit a decrease in Stokes shifts from ∼280 6).
to ∼80 cm−1 and from ∼280 to ∼120 cm−1, respectively, by the On the basis of these time-resolved spectroscopic measure-
addition of MSA (Table 4). For PF30hp with TFA, a similar ments, the S1- and T1-state lifetimes of PF30hp and PF38Np can
decrease of Stokes shift from ∼330 to ∼100 cm−1 is observed be argued to be significantly longer than those of their neutral

Table 4. Steady-State Absorption and Fluorescence Bands, Stokes Shifts, Singlet and Triplet Excited-State Lifetimes, TPA Cross
Sections σ(2), and NICS Values of Neutral and Protonated PF30h and PF38N in CH2Cl267

λabs (nm) λflu (nm) EStokes (cm−1) τS (ps) τT (μs) σ(2)(GM)(λex/nm) NICS (ppm)
PF30h 634, 1106 1141 280 47 3.1 1350 ± 200 (1290) −8.7
PF30hp 633, 1037 1046 80 400 6.1 6300 ± 500 (1300) −14.3
PF38N 725, 1280 1328 280 18 0.2 1300 ± 200 (1390) −8.7
PF38Np 746, 1260 1280 120 44 6.2 6040 ± 500 (1450) −11.5

2274 DOI: 10.1021/acs.chemrev.6b00313


Chem. Rev. 2017, 117, 2257−2312
Chemical Reviews Review

Figure 28. Femtosecond TA spectra and decay profiles (inset) of (a) Figure 29. Femtosecond TA spectra and decay profiles (inset) of (a)
neutral PF30h and (b) its protonated form PF30hp with MSA in neutral PF38N and (b) its protonated form PF38 Np with MSA in
CH2Cl2. For all cases, the pump excitation is at 630 nm. Reprinted with CH2Cl2. For all cases, the pump excitation is at 730 nm. Reprinted with
permission from ref 67. Copyright 2009 American Chemical Society. permission from ref 67. Copyright 2009 American Chemical Society.

Table 5. Fitted Time Constants in the TA Decay Profiles of Table 6. Steady-State Absorption and Fluorescence Bands,
Heptaphyrins and Nonaphyrins67 Stokes Shifts, Singlet and Triplet Excited-State Lifetimes, and
TPA Cross Sections σ(2) of PF30h and Its Protonated Form
λpump λprobe PF30hp with TFA in CHCl367
(nm) (nm) τ1 (ps) (fraction) τ2 (ps) (fraction)
PF30h 630 633, 43, 50 λabs λflu ΔEStokes τS τT σ(2) (GM)
680 (nm) (nm) (cm−1) (ps) (μs) (λex/nm)
PF30hp 630 630, 403, 396 PF30h 634, 1142 330 51 1290 ± 200
710 1101 (1290)
PF38N 730 726, 16, 220 PF30hp 633, 1050 100 167 2.0 5680 ± 500
780 1037 (1300)
PF38Np 730 726, 46, 41
780
PF30h 630 638, 48 (97%), 54 >1 ns (3%), >1 ns (14%) motions in the flexible structure of neutral forms, resulting in
680 (86%) suppression of nonradiative decay channels and deceleration of
PF30hp 630 630, 169 (69%), 165 ∼1.12 ns (31%), ∼1.28 energy-relaxation processes of the excited states for protonated
675 (62%) ns (38%)
forms. Chirvony et al. provided a similar description for the
change of photophysical properties of nonplanar porphyrin
counterparts (Table 4). Considering a great change in the S1/T1- dication (diprotonated forms of H2TPP and H2OEP).86 In their
state lifetimes and Stokes shift between neutral and protonated study, porphyrin dications show larger Stokes shifts, broadened
forms of PF30h and PF38N, these significant changes delineate optical bands, and a decrease in fluorescence lifetimes compared
structural rigidification by protonation. Because the connection to those observed in their planar neutral forms; the same spectral
of pyrrole rings by meso-methine carbons of expanded features are observed for twisted PF30h and PF38N. Chirvony et
porphyrins provides flexible structures inherently, their excited al.86 described that the altered properties of porphyrin dications
state dynamics is governed by effective nonradiative decay are attributed to their conformational flexibility and non-
processes through several low-frequency vibrational modes, and planarity. Because structural flexibility can lead to various
they show relatively short singlet excited-state lifetimes conformations in the excited electronic state, Chirvony et al.86
compared to typical porphyrins. Here, the protonation of suggested that the funnel point is enhanced at which internal
expanded porphyrins induces intermolecular interactions with conversion occurs, because of the small energy gap between the
the counteranion of acid via hydrogen-bonding interactions and ground and excited states, thereby decreasing the S1-state
Coulombic repulsion between pyrrolic protons, which leads to lifetimes in porphyrin dications. Therefore, these results support
rigidification of their structures. In reality, the X-ray structure of that the conformational changes in PF30hp and PF38N by
PF30hp under TFA condition apparently illustrates its hydro- protonation trigger a change of photophysical properties.
gen-bonding interactions with counteranions of TFA.100 As described in the Introduction, the conformational change
Namely, by protonation, structural rigidification, accompanied from a distorted to a planar structure by the effect of protonation
by the planarization of PF30hp and PF38Np, restricts molecular is supported by the TPA cross section values, reflecting the π
2275 DOI: 10.1021/acs.chemrev.6b00313
Chem. Rev. 2017, 117, 2257−2312
Chemical Reviews Review

electron delocalization. For obtaining the maximum TPA cross


section values without any linear absorption artifacts, the two-
photon excitation wavelengths were examined in the doubled
wavelength region of B-like absorption bands for neutral and
protonated species to exclude the possibility of one-photon
absorption. PF30h and PF38N show the TPA cross section
values of 1350 and 1300 GM, respectively. On the other hand,
significantly higher TPA cross section values are measured for the
protonated forms of PF30hp and PF38Np with MSA (6300 and
6040 GM, respectively). Moreover, by the addition of TFA, the
TPA cross section value of PF30h increases from 1290 to 5680
GM (Table 6). These increased TPA cross section values of the
protonated forms qualitatively illustrate that π electrons are more
effectively delocalized with the enhancement of structural
planarity induced by protonation. Nevertheless, further in-
depth studies are required for deep understanding of the
increased TPA cross section values of PF30hp and PF38Np. On
the basis of the second-order perturbation theory,102 if the
incident photon frequency is adjusted close to the energy of the
imaginary state or the transition dipole moment increases, the
TPA cross section values increase.103−106 When the TPA values
Figure 30. Calculated (a) oscillator strength and (b) NICS(0) values (in
are measured near B-like bands, the Q-like bands of expanded ppm) of neutral (top in panel a, left in panel b) and protonated (bottom
porphyrins play an essential part because they can act as an in panel a, right in panel b) forms of PF30h based on the optimized
intermediate state in the TPA process.107 The Q-like bands of the structures (B3LYP/6-31G* level). To show clearly, meso-pentafluor-
lowest energy in PF30h and PF38N become enhanced by the ophenyl rings are substituted for hydrogen atoms after the optimization
addition of acid, exhibiting blue-shifts (Figure 26). Therefore, procedures. Reprinted with permission from ref 67. Copyright 2009
strong Q-like bands lead to increased transition dipole moment American Chemical Society.
and TPA cross section values. This increase in the TPA cross
section values arises from the structural planarity and electronic
structure of the molecule.
The NICS calculation provides a quantitative analysis for the
aromaticity change by protonation. The optimized geometries of
PF30hp and PF38N for NICS calculation are relatively planar
and figure-eight distorted conformation, respectively, being
consistent with their X-ray structures (Figures 30 and 31). On
the other hand, owing to the absence of X-ray crystallographic
results, the optimized structures of PF30h and PF38Np based on
their 1H NMR spectra display figure-eight distorted and relatively
planar conformations, respectively.
In particular, five pyrrole rings are oriented inward and four
pyrrole rings are oriented outward in the optimized structure of
PF38Np, which is consistent with 1H−1H-COSY and 1H NMR
spectroscopy.67 As the calculated chemical shifts of the inner or
outer NH and β-CH protons for these optimized structures are in
agreement with the experimental chemical shifts,67 these
optimized structures are likely to be close to the actual molecular
structures, although the optimized structures were not obtained
from X-ray structures. Moreover, the optimized geometries of
PF30hp and PF38N are in agreement with the planar and figure- Figure 31. Calculated (a) oscillator strength and (b) NICS(0) values (in
eight structures, respectively, as confirmed by X-ray crystallog- ppm) of neutral (top in panel a, left in panel b) and protonated (bottom
raphy.67 In these optimized geometries, although π electrons on in panel a, right in panel b) forms of PF38N based on the optimized
the π-conjugation pathway of PF30h and PF38N follow the [4n structures (B3LYP/6-31G* level). To show clearly, meso-pentafluor-
+2] rule of Hückel aromaticity, the calculated NICS(0) values are ophenyl rings are substituted for hydrogen atoms after the optimization
in the range of small negative values (−3.1 to −7.8 ppm for procedures. Reprinted with permission from ref 67. Copyright 2009
PF30h and −0.4 to −7.1 ppm for PF38N), describing their weak American Chemical Society.
aromaticity, attributed to the largely twisted conformations. On
the other hand, along with the protonation effect giving rise to molecular center, it is sufficient for verifying the reinforced ring
the planar structures and intensified ring currents of PF30h and current by protonation, because the different distances between
PF38N, the NICS(0) values estimated in the protonated the center and π-conjugation pathway can provide a significant
structures are highly negative, −25.5 to −26.7 ppm (−14.3 effect on the NICS values. In particular, the molecular centers of
ppm in the center of macrocycle) for PF30hp and −18.8 to PF30h and PF38N are slightly close to the molecular frame
−21.9 ppm (−11.5 ppm in the cavity center of macrocycle) for because of their figure-eight structures, compared to those in
PF38Np. Although NICS values show a small increase at the their planar protonated structures of PF30hp and PF38Np. This
2276 DOI: 10.1021/acs.chemrev.6b00313
Chem. Rev. 2017, 117, 2257−2312
Chemical Reviews Review

Chart 4. Molecular Structure of PF36O (Left) and PF38O (Right) (Reprinted with Permission from Ref 68; Copyright 2010
American Chemical Society)

observation is in accordance with the NICS values around meso- Mö bius topology. For elucidating this critical point, the
positions; because the distances between the probed near-meso- protonation effect on a couple of redox [4n] and [4n+2]π
positions and molecular frame are constant regardless of the congeners, [36] and [38]octaphyrins, is comparatively discussed
conformational change, the increased NICS values at those here.68 For clarity, octaphyrins having the same meso-aryl
positions obviously indicate ring-current enhancement. In substituents, being named as meso-pentafluorophenyl-substi-
accordance with the 1H NMR spectra, the NICS values estimated tuted [36]- (PF36O) and [38]octaphyrins(1.1.1.1.1.1.1.1)
at various positions of neutral and protonated forms of PF30h (PF38O), are discussed.108,109 These compounds differ only in
and PF38N are well-matched with the fact that the aromaticity their arrangement of intramolecular hydrogen bondings between
increases through planar structures by protonation; on the other pyrroles and the number of π electrons. As the octaphyrin
hand, they display weak aromatic character in the neutral state macrocycle is sufficiently large for disregarding conformational
because of their figure-eight twisted conformations. restraint, the π electrons on the core cyclic conjugation pathway
As mentioned above, the photophysical properties of PF30h play the most important role in determining its conformation
and PF38N clearly indicate the protonation-induced conforma- under acidic conditions. By considering all factors, such large
tion changes from twisted to planar structures. Upon redox congeners in their protonated forms are most suitable for
protonation, the spectroscopic features are remarkably changed, addressing the issue of [4n+2] or [4n]π electron aromatic
and all experimental data are in agreement with the increased stabilization with respect to the geometry of expanded
planarity and rigidity; for example, relatively narrow absorption porphyrins. Here, the discussion about the relationship between
bands, decreased Stokes shift, significantly higher TPA cross aromaticity and photophysical properties can offer the deep
section values, and longer excited-state lifetimes of PF30hp and comprehension of Hückel or Möbius aromatic switching by the
PF38Np are observed as compared with their figure-eight protonation effect.
twisted neutral forms. Consequently, more planar structures give Through the X-ray crystallography, a figure-eight conforma-
rise to enhanced aromatic character, resulting in highly negative tion of PF36O being composed of two porphyrin-like
NICS values. These results indicate that the conformational tetrapyrrolic “hemimacrocycles” is clearly revealed (Chart 4).68
distortion of expanded porphyrins, which is inevitable partic- All pyrrolic nitrogen atoms are oriented inward, constructing an
ularly with the increasing number of pyrroles, can be modified intramolecular hydrogen-bonding network between the pyrrole
and controlled by protonation, as well as anion binding leading to constituents. PF36O shows proton peaks in the 6−8 ppm region
planar structures to enhance the Hückel aromaticity. in the 1H NMR spectrum. Moreover, the NICS values in the
3.1.2. Effects of Protonation in [4n]π Expanded range from −2.1 to 2.1 ppm are estimated at several positions
Porphyrins. Under neutral conditions, the intramolecular around the macrocycle.68 These results illustrate that PF36O is
hydrogen bonding of expanded porphyrins leads to a large nonaromatic.110 In the steady-state absorption spectrum of
difference in their conformation and the number of π electrons PF36O, two broad bands in the visible region without any Q-like
participating in core cyclic π-conjugation. For example, in bands (Figure 32) are also characteristic spectral features of
nonpolar solvents, PF30h and PF38N exhibit figure-eight nonaromatic expanded porphyrins.
structures, despites their weak [4n+2]π aromaticity, because Even though there is no crystal structure of PF38O, its 1H
the more effective intramolecular hydrogen bondings are formed NMR spectrum illustrates the diatropic ring current. The well-
in the figure-eight conformation.66 On the other hand, under resolved β-proton signals in the moderately shielded region (δ =
acidic conditions, protonation breaks the intramolecular hydro- 2.51 and 2.78 ppm) arise from the inner pyrrolic protons Ha and
gen bonds between pyrroles in expanded porphyrins. Instead, the Hb in Chart 4, and other signals in the deshielded region result
released aminic pyrrole units in expanded porphyrins interact from the outer pyrrolic protons, which are indicative of its
with counteranions of acid molecules and form intermolecular obvious diatropic ring current.68 Moreover, as the low-temper-
hydrogen bonds, often resulting in drastic transformation of their ature 1H NMR analysis of PF38O at 183 K did not exhibit any
conformation, in addition to the incresed aromaticity. Generally, broadening of signal, these proton signals represent the less-
the above-mentioned PF30h and PF38N achieve the [4n+2]π planar Hückel aromatic conformation,62 in which the protons are
Hückel planar structures under acidic conditions, resulting from not entirely lying in the shielding region of the diatropic ring
the aromatic stabilization of planar topology. Nevertheless, it is current. In addition, the total eight signals for 16 protons reflect
not totally excluded by this result that the planar form is not the substantially symmetric structure in solution on the 1H NMR
most stable conformation for [4n]π expanded porphyrins timescale. In the steady-state absorption spectrum of PF38O, a
because [4n]π electronic systems can undergo topological B-like band in the visible region and Q-like bands in the NIR
changes through aromatic stabilization resulting in twisted region are observed, which are typical spectral features of
2277 DOI: 10.1021/acs.chemrev.6b00313
Chem. Rev. 2017, 117, 2257−2312
Chemical Reviews Review

analysis, the protonated species is assigned to be a dication of


PF36O (PF36O-TFA2),68 where a large, flexible core macro-
cycle of [36]octaphyrin permits smooth π-conjugation on an
overall half-twisted Möbius topology, where torsional angles
along the π-conjugation circuit are well-settled within 30°.30,33
The temperature-dependent 1H NMR spectra of [36]-
octaphyrins with controlling the addition of acid amount provide
the change of aromaticity and conformation in detail.68 In the
TFA titration NMR spectra in CD3CN at 233 K, the signal
intensities from protonated species, such as PF36O-TFA1 and
PF36O-TFA2, are increased at the expense of the decrease in the
signal intensities from PF36O. Upon the addition of 3 equiv of
TFA, the signals of PF36O disappear and new signals from
PF36O-TFA1 and PF36O-TFA2 simultaneously appear. Several
proton signals in the region of −1 to −5 ppm indicate the large
shielding effect on inner protons, which reflects their aromaticity.
With the further addition of 10 equiv of TFA, a simple 1H NMR
spectrum consisting of signals of only PF36O-TFA2 is obtained.
Furthermore, with even an excess amount of TFA or its deutrated
form, there is no spectral change, indicating no further
protonated species, such as PF36O-TFAn (n = 3 or 4). Thus,
TFA only allows the protonation of PF36O up to a dication
species. The titration with MSA exhibits similar changes in the
absorption spectra, which is indicative of the production of
Figure 32. Steady-state absorption (black) and fluorescence spectra PF36O-MSA1 and PF36O-MSA2 with a mild counteranion
(red) of PF36O, PF38O, and their protonated forms (PF36O-MSA2, effect in the protonation of PF36O. The highly shielded inner
PF36O-MSAn, and PF38O-MSA2, respectively) upon addition of MSA protons represent that these protonated species exhibit strong
in CH2Cl2. Reprinted with permission from ref 68. Copyright 2010 aromaticity and that subtle enhancement is observed from
American Chemical Society. PF36O-MSA1 to PF36O-MSA2. Because of the stronger acidity
of MSA, the proton signals of PF36O almost disappear by the
aromatic expanded porphyrins (Figure 32).77,79,94 These spectral addition of equivalent MSA.
features of neutral PF38O illustrate a [38]π electronic Hückel In addition, 4 equiv of MSA results in new signals, in addition
aromatic conformation with two inverted pyrrole rings, as shown to those of PF36O-MSA2, which are probably induced by
in Chart 4. PF36O-MSAn (n = 3 or 4). The moderate chemical shifts of the
Figure 33 shows the crystal structure of protonated PF36O.68 shielded inner protons propose that, as compared to PF36O-
By protonation, the figure-eight geometry of neutral PF36O is MSA1 and PF36O-MSA2, PF36O-MSAn exhibits relatively weak
changed to an unfolded conformation. All intramolecular aromaticity.68
hydrogen bondings in PF36O are broken and changed to The acid titration to PF36O with MSA gives rise to the two-
intermolecular hydrogen bondings between pyrrolic protons and stepwise absorption spectral changes.68 Up to the amount of 2.72
TFA molecules by protonation, in which five nitrogens atoms × 10−4 M MSA addition, a distinct spectral change occurs with
(located at the pyrrole rings marked as A, C, D, F, and G in Figure two isosbestic points at 426 and 667 nm, respectively. With
33) are oriented outward for forming hydrogen bonds with increasing intensity, a broad band in the visible region is red-
surrounding TFA molecules. From X-ray crystallographic shifted, and additional broad and weak bands appear in the NIR

Figure 33. X-ray crystal structure of diprotonated PF36O-TFA2. meso-Aryl substituents are omitted for clarity. Reprinted with permission from ref 68.
Copyright 2010 American Chemical Society.

2278 DOI: 10.1021/acs.chemrev.6b00313


Chem. Rev. 2017, 117, 2257−2312
Chemical Reviews Review

range. These spectral features, sharp B-like and distinct Q-like


bands, are typically characteristic of aromatic porphyri-
noids.77,79,94 Hence, this spectral change illustrates the enhance-
ment of aromaticity by protonation. With increasing MSA
concentration up to 3.79 × 10−3 M, the B-like band becomes
more intensified and the Q-like bands become well-resolved and
intense. Because of the complicated titration process of PF36O
by protons, it is difficult to obviously distinguish each species
from various protonated forms. Nevertheless, the titration results
analyzed by Hill plots describe that two protons are attached in
the first titration process using an MSA concentration of 2.72 ×
10−4 M (Figure 34).

Figure 35. Analyses of the protonation process of PF36O: (a) analyzed


absorption spectra of three protonated forms and (b) progress of
protonation from neutral to mono-, di-, and tetraprotonated forms
(inset describes concentration range from 0 to 0.0008 M). Reprinted
with permission from ref 68. Copyright 2010 American Chemical
Society.

protonated species of PF36O in spectrophotometric titration up


to MSA concentrations of 2.72 × 10−4 and 3.79 × 10−3 M can be
assigned to diprotonated PF36O-MSA2 and multiprotonated
PF36O-MSAn, respectively (Figures 32 and 36).
Figure 37 displays the crystal structure of protonated
PF38O.68 The structure of protonated PF38O is surprisingly
similar to that of protonated PF36O (PF36O-TFA2). The
octaphyrin macrocycle exhibits a mirror symmetry with a plane
being vertical to A and E pyrrole rings. Five nitrogen atoms
(located at the pyrrole rings marked as A, C, D, F, and G in Figure
37) are oriented outward as those of PF36O-TFA2. The
octaphyrin molecules form intermolecular hydrogen bonds with
surrounding TFA, ethanol, isopropyl alcohol, and water
molecules. The protonated species is considered again to be
Figure 34. Hill plots for the titration of PF36O in an MSA the dication form of PF38O (PF38O-TFA2) based on their
concentration range of (a) 0 to 1.09 × 10−5 M and (b) 1.09 × 10−5 to
2.72 × 10−4 M. Reprinted with permission from ref 68. Copyright 2010
number of counteranions. Despite the fact that PF36O-TFA2
American Chemical Society. and PF38O-TFA2 are structurally similar, the π-conjugation
circuit of PF38O-TFA2 exhibits Hückel planar geometry.
Notably, the distinction of Möbius−Hückel geometry between
The singular-value decomposition (SVD) analysis for the PF36O-TFA2 and PF38O-TFA2 is attributed to the small
spectrophotometric titration spectra provides three linearly difference in the dihedral angles between F and G pyrrole rings.
independent optical spectra assigned to arise from monoproto- In the crystal structure of PF38O-TFA2, all pyrrole planes
nated, diprotonated, and tetraprotonated species, respectively including F and G faces are directed to almost the same direction,
(Figure 35).68 At an MSA concentration of <0.6 × 10−3 M, the affording an approximately Hückel planar structure, while F and
spectrum of the monoprotonated form disappears, while those of G rings in PF36O-TFA2 are substantially distorted to the plane
di- and tetraprotonated species are observed in a wide composed of the other pyrrole rings and show overall Möbius
concentration range of 1.30 × 10−4 to 5.8 × 10−3 M and 5.8 × conformation in the X-ray structure of PF36O-TFA2.
10−3 to 0.8 × 10−2 M, respectively. Namely, the di- and With the addition of TFA or MSA, a one-stepwise simple
tetraprotonated species exist as the two major species in the spectral change, from PF38O to the protonated forms, occurs in
whole protonation process with MSA. Consequently, the two the 1H NMR titration analysis. Even though the protonated form
2279 DOI: 10.1021/acs.chemrev.6b00313
Chem. Rev. 2017, 117, 2257−2312
Chemical Reviews Review

Figure 36. Spectrophotometric titration of (a) PF36O and (b) PF38O with MSA in CH2Cl2. Reprinted with permission from ref 68. Copyright 2010
American Chemical Society.

Figure 37. X-ray crystal structure of PF38O-TFA2. meso-Aryl substituents and hydrogen atoms of solvent molecules are omitted for clarity. Reprinted
with permission from ref 68. Copyright 2010 American Chemical Society.

shows very broad signals even at 183 K, this suggests that the large chemical shift in PF38O-MSA2 probably reflects that its
protonated species clearly possesses a symmetric conforma- conformation becomes more planar by diprotonation. Upon the
tion.68 So, the protonated species possibly corresponds to the addition of MSA up to a concentration of 5.14 × 10−5 M, even
diprotonated species of [38]octaphyrin (PF38O-TFA2 and though there is no significant spectral change, the absorption
PF38O-MSA2); its structure exhibits mirror symmetry as spectrum indicates an almost one-step spectral change with two
revealed by X-ray crystallography. Signals for monoprotonated isosbestic points at 545 and 1009 nm, respectively (Figure 36).
forms (PF38O-TFA1 and PF38O-MSA1) are not observed in Further addition of MSA gives rise to no spectral change. Hence,
titration experiments. In the spectrum of PF38O-MSA2, an the absorption spectrum at 5.14 × 10−5 M of MSA condition
increased shielding effect is observed by protonation. Fur- represents that of PF38O-MSA2, in which the protonation
thermore, the most shielded signal of PF38O-MSA2 appears near results in intensified and red-shifted B- and Q-like bands. The
−10 ppm, while that of PF38O is detected only at 2.5 ppm; this intensified B- and Q-like bands indicate the enhancement of
2280 DOI: 10.1021/acs.chemrev.6b00313
Chem. Rev. 2017, 117, 2257−2312
Chemical Reviews Review

Figure 38. Femtosecond TA decay profiles and spectra of neutral (PF36O) and protonated forms (PF36O-MSA2 and PF36O-MSAn) of
[36]octaphyrin. For all cases, the pump excitation is at 620 nm. Reprinted with permission from ref 68. Copyright 2010 American Chemical Society.

Figure 39. Femtosecond TA decay profiles and spectra of neutral (PF38O) and protonated forms (PF38O-MSA2) of [38]octaphyrin. For all cases, the
pump excitation is at 620 nm. Reprinted with permission from ref 68. Copyright 2010 American Chemical Society.

aromaticity along with the protonation from PF38O to PF38O- For obtaining in-depth information on the protonation effect
MSA2, being well-matched with 1H NMR analysis. On the basis on the photophysical properties of expanded porphyrins, the
of 36 and 38π electrons of PF36O-TFA2 and PF38O-TFA2, excited-state dynamics of the neutral and protonated octaphyrins
respectively, protonated octaphyrins prefer to undergo con- are comparatively investigated.48 In the fs-TA spectra of neutral
formational change to twisted and planar conformation, and protonated [36]octaphyrins, the decay profiles of GSB
respectively, for realizing [4n]/[4n+2] Möbius/Hückel aroma- recovery and ESA signals display complicated decay dynamics.
ticity to fulfill the rule determined by the number of π electrons. The lifetime constant of <1 ps can be ascribed to an energy
This facile conformational change infers that the size of the relaxation process from the high to the lowest excited state, such
octaphyrin macrocycle is large enough for preferred topology as a relaxation process from the Sn- to S1-state.112 As PF36O
control without severe structural constraints. Under acidic possesses conformational flexibility, its TA spectra display probe-
conditions, protonation eliminates intramolecular hydrogen- wavelength dependences on the S1-state lifetimes (Figure
bonding interactions, which allows these octaphyrin macrocycles 38).112−119 On the other hand, PF38O does not show any
to transform to the most stable conformation determined by probe-wavelength dependence in the S1-state lifetime (Figure
aromatic stabilization effect depending on their number of π 39). Neither the GSB nor ESA signal of PF36O is observed in
electrons, protonated PF36O and PF38O for Hückel planar and nanosecond TA measurements, owing to its rapid internal
Möbius twisted topology.111 conversion process and less efficient intersystem crossing, being
2281 DOI: 10.1021/acs.chemrev.6b00313
Chem. Rev. 2017, 117, 2257−2312
Chemical Reviews Review

Table 7. Steady-State Absorption and Fluorescence Bands Maxima, Stokes Shifts, Singlet and Triplet Excited-State Lifetimes, TPA
Cross Section σ(2), and NICS and BLA Values68
molecules λabs (nm) λflu (nm) ΔEStokes (cm−1) τS (ps) τT (ps) σ(2) (GM) (λex/nm) NICS (ppm) ΔRx (Å)
[36]octaphyrin PF36O 637 8.6 800 (1280) −2.1 0.126
PF36O-MSA2 731, 791, 1250 1275 157 32.1 18.3 5100 (1450) −9.3 0.083
PF36O-MSAn 742, 1228 1244 105 41.6 0.9, 8.8 2500 (1460)
[38]octaphyrin PF38O 719, 1115 1180 494 12.1 1.38 1800 (1420)
PF38O-MSA2 746, 1227 1240 111 40.7 0.22, 2.90 4600 (1420) −8.8 0.068

Figure 40. Open-aperture femtosecond Z-scan traces of (a) PF36O and (b) PF38O upon protonation forms. The red lines are the best-fit curves for the
experimental data. Reprinted with permission from ref 68. Copyright 2010 American Chemical Society.

in a good accordance with the rapid internal conversion its twisted conformation. On the basis of the results by 1H NMR
processes of the singlet excited state. as well as the steady-state and time-resolved spectroscopy,
On the other hand, weak triplet ESA and GSB recovery signals PF38O contains Hückel aromatic structures. So, compared to
with a lifetime of 1.38 μs are observed for PF38O,68 supporting PF36O, PF38O exhibits a higher TPA cross section value of
that the planar Hückel aromatic structure of PF38O is preserved 1800 GM. For protonated PF36O and PF38O, the TPA cross
because of [4n+2]π conjugated electrons, even in its neutral section values are greater than those of neutral species because of
species. After the addition of MSA, the S1-state lifetimes of effective π-delocalization and rigid conformation of protonated
PF36O-MSA2 (32 ps) and PF36O-MSAn (42 ps) are obviously species; PF36O-MSA2 and PF38O-MSA2 exhibit the TPA cross
elongated versus that of PF36O (Figure 38). Similarly, the decay section values of 4600−5100 GM, greater than those of their
profiles of GSB recovery and ESA signals of PF38O-MSA2 show neutral congeners (Table 7).
slower decay time constants (Figure 39). In addition, intense In addition, quantum mechanical calculations describe the
triplet ESA and GSB recovery signals are observed in PF36O- enhancement of the aromaticity of expanded porphyrins by
MSA2 (18.3 μs), PF36O-MSAn (0.9 and 8.8 μs), and PF38O- protonation.68 The NICS values of neutral and protonated
MSA2 (0.22 and 2.90 μs).68 Although protonated octaphyrins octaphyrins illustrate the increased aromatic character in the
PF36O-MSA2 and PF36O-MSAn exhibit complex decay protonated forms. The NICS values of PF36O calculated at
dynamics, which is fitted with two time constants, the triplet- various positions range from −2.1 to 2.1 ppm, where these low
state quantum yield is clearly enhanced by the addition of MSA NICS values are attributed to its distorted conformation and
(Table 7). nonaromatic character based on its [4n]π electrons.110 On the
The aromaticity change in octaphyrins by protonation is other hand, the NICS value of −9.3 ppm is calculated for the
described by the TPA cross section values based on their π optimized geometry of PF36O-TFA2 from its X-ray structure.
electron conjugation (Figure 40).77,79,94,113,114 PF36O exhibits a So, the change of NICS values between neutral and protonated
relatively small TPA cross section value of 800 GM because its π- PF36O well-describe that the aromatic character of PF36O-
conjugation over the whole molecular framework is disrupted by TFA2 is intensified because of its Möbius conformation by
2282 DOI: 10.1021/acs.chemrev.6b00313
Chem. Rev. 2017, 117, 2257−2312
Chemical Reviews Review

protonation. Because identical bond lengths in macrocycles are photophysical properties, which can influence their excited-state
induced by the efficient π-conjugation pathway, BLA calculation dynamics and aromaticity; thus, deuterium exchange121−124 and
is one of the effective methods for estimating molecular deprotonation experiments70,125,126 have been fulfilled for
aromaticity.120 The BLA value of 0.126 Å for PF36O is revealing the effect of deprotonation on their aromaticity.
decreased to be 0.083 Å for PF36O-TFA2 (Table 7). This PF26H108 has been well-known for a typical expanded
suggests that the overall geometric changes of expanded porphyrin having a 26π-conjugation pathway in its planar
porphyrins by protonation give rise to intense aromatic structure to comply with the Hückel [4n+2] rule (Chart 5). By
character. Consequently, protonated expanded porphyrins,
especially PF36O-TFA2 (or PF36O-MSA2), adopt the Möbius Chart 5. Molecular Structures and π-Conjugation Pathways of
structure with a distinct aromatic feature. PF26H, PF26H−, and PF26H2−; meso-Aryl Groups and Outer
Protonated structures and aromatic characters are unambig- Hydrogen Atoms Are Omitted for Clarity in Side Views of
uously discussed from their 1H NMR spectra, steady-state Each Optimized Structure; Ar = 2,3,4,5,6-Pentafluorophenyl
absorption, and X-ray crystallography analyses. By protonation, (Reprinted with Permission from ref 70; Copyright 2012
nonaromatic PF36O with figure-eight conformation changes to Wiley-VCH Verlag GmbH & Co. KgaA)
strong Möbius aromatic forms: PF36O-TFA1 and PF36O-TFA2
(or PF36O-MSA1 and PF36O-MSA2). The crystal structure of
PF36O-TFA2 illustrates that effective intermolecular hydrogen
bondings give rise to the widely extended conformation, where
the hydrogen-bonding network is composed of the macrocycle,
counteranions, and acid molecules by breaking the intra-
molecular hydrogen bonds of neutral species by protonation.
On the other hand, by protonation, weak Hückel aromatic
PF38O is changed to the strong Hückel aromatic form of
PF38O-TFA2 (or PF38O-MSA2). The crystal structure of
PF38O-TFA2 exhibits a conformation surprisingly similar to that
of PF36O-TFA2, although the Hückel topology of PF38O-TFA2
is clearly distinguished from the Möbius topology of PF36O-
TFA2. On the basis of these results, it is strongly suggested that,
under acidic conditions, where the macrocycle exhibits
conformation free from intramolecular hydrogen bonds, large
expanded porphyrins adopt their aromatic topology between [4n
+2] Hückel and [4n] Möbius conformations according to their
number of π electrons with a gain of aromatic stabilization.
Generally, as [36]octaphyrins assume figure-eight conformations
due to the flexibilities of molecular frameworks, [36]octaphyrins
exhibit (1) fast singlet excited-state dynamics and low triplet state
quantum yield, (2) no Q-like bands, (3) low TPA cross section the elaborate treatment of PF26H with adding 1 equiv of TBAF,
values, and (4) no fluorescence. On the other hand, by PF26H− monoanion is formed, and the addition of 10 equiv of
protonation, because the protonated PF36O and PF38O have TBAF exclusively affords PF26H2− dianion.67 The formation of
rigid Möbius or Hückel aromatic conformations by intermo- PF26H− and PF26H2− is observed by their 1H NMR spectra.
lecular hydrogen-bonding networks, the following are man- The structure of PF26H2− is unambiguously revealed by X-ray
ifested: (1) a Q-like band in the NIR region, (2) a red-shifted crystallography.127 These three hexaphyrins, PF26H, PF26H−,
sharp B-like band, (3) increased TPA cross section values, and and PF26H2−, commonly show aromatic features, but differ-
(4) increased excited-state lifetime. These experimental results ences in 1H NMR peak shifts between the inner and outer
provide an opportunity for demonstrating a close relationship protons (Δδ) increase going from PF26H to PF26H− and
among structures and photophysical properties, such as PF26H2−, implying an enhanced diatropic ring current in this
calculated and observed molecular features, and excited-state sequence (Figure 41).
dynamics. Figure 42 shows the absorption and fluorescence spectra of
PF26H, PF26H−, and PF26H2−. Going from PF26H to
3.2. Control of Aromaticity by Deprotonation
PF26H− and PF26H2−, the intensities of B- and Q-like bands
3.2.1. Effects of Deprotonation in [4n+2]π Expanded increase, and the whole absorption spectra become more
Porphyrins. Protonation induces the extension of molecular resolved. Importantly, the fluorescence of PF26H93 in the NIR
structures of expanded porphyrins by breaking the hydrogen region is considerably magnified around 20 and 36 times in
bonds of their inner cavity. Deprotonation is also known as one PF26H− and PF26H2−, respectively. These intensified fluo-
of the effective methods for the changes of conformation and rescence spectra are remarkable considering the recent growing
aromaticity in expanded porphyrins because the elimination of interest in NIR dyes required in biosensors, solar cells, and
pyrrolic protons by deprotonation also breaks the hydrogen optical communication.128−132 Initially, these intensified fluo-
bonds of their inner cavity. Herein, as a part of continuous efforts rescence spectra are understood with regard to the inhibition of
on controlling the aromaticity of expanded porphyrins, a method NH tautomerization and increased aromaticity in PF26H− and
to control the aromaticity of [26]hexaphyrin by deprotonation PF26H2− based on their absorption spectra.
using tetrabutylammonium fluoride (TBAF) is described. In A deuterium-exchange experiment for the inner NH protons
expanded porphyrins, intramolecular hydrogen bonds have been of PF26H with deuterium clarifies the origin of change in the
identified to cut a brilliant figure in their molecular structures and absorption spectra. Nevertheless, there are no distinct features in
2283 DOI: 10.1021/acs.chemrev.6b00313
Chem. Rev. 2017, 117, 2257−2312
Chemical Reviews Review

Figure 41. 1H NMR spectra of (i) PF26H− and (ii) PF26H2− in CDCl3. Residual solvents, x; residual peaks of PF26H2−, X. Reprinted with permission
from ref 70. Copyright 2012 Wiley-VCH Verlag GmbH & Co. KgaA.

PF26H. As similar deprotonation experiments are performed for


pentafluorophenyl porphyrin in dimethylsulfoxide (DMSO), the
fluorescence quantum yield decreases going from freebase to
dianion; hence, the intensified fluorescence spectra of the
deprotonated species PF26H− and PF26H2− are attributed to
enhanced aromaticity.94
The deprotonation results of PF26H are clearly supported by
its acidic nature, which is well-matched with the observation that
PF26H spontaneously forms PF26H− in hydrogen-bonding
acceptor solvents, such as DMSO and dimethylformamide
(DMF) (Figure 43). By spectrophotometric titration, the
equilibrium constants are estimated to be K1 = 2.5 × 1010
between PF26H and PF26H− and K2 = 5.3 × 108 between
PF26H− and PF26H2−.70,133 More clearly, the acidic nature of
Figure 42. Steady-state absorption (black line) and fluorescence (gray PF26H can be revealed by the evaluation of its ΔpKa value of 2.6
line) spectra of PF26H, PF26H−, and PF26H2− in CH2Cl2. Reprinted with regard to 1,8-diazabicyclo-[5,4,0]undec-7ene (DBU).134
with permission from ref 70. Copyright 2012 Wiley-VCH Verlag GmbH The basicity of DBU is not strong enough to affect the
& Co. KgaA.
deprotonation of H2TPP at its highest concentration. However,
DBU can deprotonate PF26H− as well as PF26H, and under the
the shape and intensity of fluorescence spectra during the same conditions, the titration experiment permits the determi-
deuterium-exchange experiment, indicating that NH tautomeri- nation of ΔpKa value of 8.3 for PF26H−, obviously representing
zation does not substantially affect the nonradiative decay rate of the acidic nature of PF26H.

Figure 43. Normalized absorption (solid lines) and fluorescence (dashed lines) spectra of PF26H in various solvents. Reprinted with permission from
ref 70. Copyright 2012 Wiley-VCH Verlag GmbH & Co. KgaA.

2284 DOI: 10.1021/acs.chemrev.6b00313


Chem. Rev. 2017, 117, 2257−2312
Chemical Reviews Review

Figure 44. TA spectra (top) and decay profiles (bottom) of PF26H− (left) and PF26H2− (right) in toluene. Reprinted with permission from ref 70.
Copyright 2012 Wiley-VCH Verlag GmbH & Co. KgaA.

Figure 45. (a) ACID plots of PF26H, PF26H−, and PF26H2− with an isosurface value of 0.08; (b) NICS values for PF26H, PF26H−, and PF26H2−;
and (c) 26-, 25-, and 24-membered conjugated π-electronic circuits of PF26H, PF26H−, and PF26H2−, respectively. Reprinted with permission from ref
70. Copyright 2012 Wiley-VCH Verlag GmbH & Co. KgaA.

By femtosecond TA spectroscopy, the S1-state lifetime of respectively (Figure 44). The long-lived S1-states and intensified
PF26H is estimated to be ∼100 ps in toluene, while those of emissions of PF26H− and PF26H2− as compared to PF26H
PF26H− and PF26H2− are much longer, i.e., 740 and 870 ps, indicate that the nonradiative decay rates of the S1-states for
2285 DOI: 10.1021/acs.chemrev.6b00313
Chem. Rev. 2017, 117, 2257−2312
Chemical Reviews Review

PF26H− and PF26H2− are significantly reduced with regard to a 26π-electronic circuit having two outer rims of the diagonal
the radiative decay rates. By nanosecond TA spectroscopy, the amino-type pyrroles and the four nitrogen atoms of the imine-
T1-state lifetimes of PF26H− and PF26H2− are measured to be type pyrroles. For this circuit, the HOMA values of PF26H and
10.4 and 22 μs, respectively; these values are also substantially PF26H2− are estimated as 0.647 and 0.571, respectively.127
longer than that of PF26H (2 μs).70 Thus, deprotonation results Considering the symmetry of PF26H2−, a different conjugation
in extended S1- and T1-state lifetimes. pathway emerges for better macrocyclic π-electronic circuit,
As described in the Introduction, the TPA cross section values which comprises a 24π-electronic circuit passing all six nitrogen
are closely correlated with the degree of aromaticity in expanded atoms and without including any of the peripheral CC bonds
porphyrins; these values also rely on the molecular silhouette and in the π-conjugation pathway (Figure 45). The HOMA value
symmetry of molecules.135,136 Namely, the aromaticity is along this circuit is calculated as 0.84, representing more effective
intuitively related to the TPA phenomena.94 The TPA cross π-conjugation along this circuit.70 In this connection, the
section values evaluated for PF26H, PF26H−, and PF26H2− are HOMO of PF26H2− displays symmetric electronic structures,
1000, 630, and 1450 GM, respectively.70 The higher TPA cross in which the electron density is localized only on this 24π-
section value for PF26H2− is considered to support the electronic circuit. Likewise, a 25π-electronic circuit has been
conjecture that PF26H2− exhibits a more efficient conjugated suggested for the more effective π-conjugation of PF26H−, as
Hückel aromatic circuit, while the smaller TPA cross section illustrated in Figure 45. Nevertheless, in these π-conjugation
value of PF26H− is possibly originated from its nonsymmetric pathways, the number of π electrons in the π-conjugation
electronic structure. The TPA cross section values are influenced pathway remains 26, with the nonbonding pair electrons of the
by the molecular symmetry in the same chemical environment as nitrogen atom; thus, PF26H− and PF26H2− also follow the
well as the degree of aromaticity. Hückel [4n+2] rule. Another structural benefit for preferable π-
As illustrated in Chart 5 and Figure 45, compared to PF26H conjugation by deprotonation is enhanced planarity; the mean
and PF26H2−, PF26H− exhibits an asymmetric molecular plane deviation (MPD), which represents the degree of
structure and electron density, which possibly results in the structural distortions of the macrocycles from planar structures,
decreased TPA cross section values. The optimized geometries decreases from PF26H (0.41) to PF26H2− (0.16) (Figure
and their vertical transitions for PF26H, PF26H−, and PF26H2− 47).70,127 Diminished steric hindrance between the inner NH
provide better insight into their electronic structures. and pyrrolic β-protons by deprotonation possibly contributes to
The four FMOs play a key role in determining the the planar structure of PF26H2−. As a different approach, the
photophysical properties of expanded porphyrins (Figure 46). ACID plots can directly plot the magnitude and direction of the
induced ring current during the application of an external
magnetic field to a molecule.137
Accordingly, this analysis is suitable for illustrating the
aromaticity and π electron delocalization. Along with the above
result, the π-electronic circuit of PF26H constitutes substantial
electron density at the outer rim of the two amine-type pyrrole
rings. On the contrary to PF26H, the π-conjugation pathways of
PF26H− and PF26H2− exhibit different routes, especially at the
parts of the deprotonated pyrroles, which is in line with the above
results, namely, the π-conjugation pathways in PF26H− and
PF26H2− flow along the nonbonding pair electrons of the
nitrogen atoms in anionic pyrrole rings, as illustrated in Figure
45. The NICS calculation quantitatively evaluates the aromaticity
of these molecules.52 PF26H, PF26H−, and PF26H2− exhibit
negative NICS values, indicating their aromatic characters
Figure 46. Frontier orbitals of PF26H, PF26H−, and PF26H2− (Figure 45). Because of different electronic structures including
obtained by time-dependent density functional theory (TD-DFT) the direction of each π-conjugation pathway and molecular
calculations at the B3LYP/6-31G(d) level. HOMO−LUMO energy topology, it is difficult to estimate accurate central positions. The
gaps (ΔE) of the three species are almost equal to 1.80 eV. Reprinted results obtained for the selected NICS(0) values in PF26H,
with permission from ref 70. Copyright 2012 Wiley-VCH Verlag GmbH PF26H−, and PF26H2− indicate that the degree of aromaticity is
& Co. KgaA. slightly reduced. The small decrease in the NICS(0) values for
PF26H− and PF26H2− as compared with that of PF26H is
Surprisingly, the energy gaps between HOMO and LUMO of possible evidence for structural planarization. The central Bq
PF26H, PF26H−, and PF26H2− are almost equal to 1.80 eV, and ghost atoms of PF26H, PF26H−, and PF26H2− are affected by
the energy gap between LUMO and LUMO+1 of PF26H2− the local aromaticity of each pyrrole as well as the global
becomes very small (0.017 eV). As compared to PF26H, the aromaticity induced by the main π-electronic circuit. With
deprotonated species PF26H− and PF26H2− show more respect to the main 26π-conjugation pathway, the central Bq
degenerate energy states between LUMO and LUMO+1, ghost atoms are located inside the global diatropic ring current.
characteristic features of general aromatic porphyrinoids, e.g., However, these Bq ghost atoms are considered to locate outside
metalated porphyrins, following Gouterman’s four-orbital the local aromaticity of the respective pyrrole ring. With
model.55 The comparison between the X-ray structures of decreasing tilting angles of N-inverted pyrrole rings, the
PF26H and PF26H2− reveals that the deprotonation process NICS(0) values are more influenced by the local aromaticity of
leads to considerable structural deformation, which apparently the respective pyrrole rather than the global aromaticity by the
affects electronic properties.70 The bold line of PF26H in Chart 5 main π-conjugation pathway, because of their spatial positions.
represents the 26π-conjugated aromatic circuit, which comprises Nonetheless, all NICS(0) values at the center of molecules
2286 DOI: 10.1021/acs.chemrev.6b00313
Chem. Rev. 2017, 117, 2257−2312
Chemical Reviews Review

Figure 47. Comparison of the MPD diagrams of PF26H, PF26H−, and PF26H2− on the basis of their optimized structures. Reprinted with permission
from ref 70. Copyright 2012 Wiley-VCH Verlag GmbH & Co. KgaA.

exhibit significantly large values; thus, PF26H, PF26H−, and whether PF26H exhibits the proton-transfer effect on the
PF26H 2− are considered as strong aromatic expanded fluorescence feature. On the other hand, the same absorption
porphyrins in accord with their 1H NMR spectra. More spectra and fluorescence intensity suggest that tautomerization
importantly, the NICS(0) values were also estimated at the does not affect the radiative decay process in [26]hexaphyrins.
center of the respective pyrrole. In PF26H, the calculated NICS Hence, deprotonation results in substantial structural deforma-
values are negative at the center of amine-type pyrroles, implying tion, which affords different and more effective π-conjugation
that those positions are significantly involved in the global pathways for PF26H− and PF26H2−, which are responsible for
aromaticity induced by macrocyclic ring. Nevertheless, the NICS reinforced aromaticity.
values at the deprotonated amine-type pyrroles of PF26H− and 3.2.2. Effects of Deprotonation in [4n]π Expanded
PF26H2− are positive, implying that the main π-conjugation Porphyrins. As explained above, the deprotonation of PF26H
pathways circumvent the nitrogen atoms in the deprotonated with TBAF results in the formation of mono- and dideproto-
amine-type pyrroles for PF26H− and PF26H2−. Taken together, nated [26]hexaphyrins, which exhibit increased degree of
all the results underpin the contention that the reinforced aromaticity and strong fluorescence.70 On the basis of these
aromaticity of PF26H− and PF26H2− can be examined by results, the deprotonation of [32]heptaphyrins is also
considering new π-conjugation pathways, 25- and 24-membered described.69 Deprotonation is effective for realizing Möbius
π-electronic circuits, respectively, caused by structural deforma- aromatic heptaphyrins.
tion due to deprotonation. Another evidence for reinforced To delineate the effects of deprotonation in expanded
aromaticity is considered with regard to enhanced fluorescence porphyrins, meso-(pentafluorophenyl)-substituted (PF32h)
intensity in deprotonated hexaphyrins. For porphycenes, and meso-(2,6-dichlorophenyl)-substituted [32]heptaphyrin
structural distortion is induced by the tautomerization, resulting (DCP32h) are chosen, and N-fused meso-(pentafluorophenyl)-
in the increase of nonradiative decay process.121−124 However, substituted [32]heptaphyrin is chosen (N-PF32h) as a
intensified fluorescence of PF26H− and PF26H2− do not seem structurally fixed molecule (Chart 6).66,138Figure 48 shows the
to originate from the suspension of possible proton transfer steady-state absorption spectra of neutral PF32h, DCP32h, and
because PF26H exhibits a slightly distorted structure relative to N-PF32h and their deprotonated species in CH2Cl2. In the case
that of porphycene. Furthermore, deuterium exchange at the of PF32h, a broad B-like band without distinct Q-like bands is
inner nitrogen with D2O in PF26H is conducted for determining observed, characteristic of nonaromatic expanded porphyrins. In
2287 DOI: 10.1021/acs.chemrev.6b00313
Chem. Rev. 2017, 117, 2257−2312
Chemical Reviews Review

Chart 6. Molecular Structures of Neutral PF32h, DCP32h, addition of an excess amount of base into DCP32h in CDCl3
and N-PF32h (Reprinted with Permission from Ref 69; results in a large aromatic ring current, as indicated by the inner
Copyright 2014 Royal Society of Chemistry) β-CH protons in highly upfield shifted and the outer β-CH
protons at the downfield region around 9 ppm. Deprotonation
also results in red-shifted B- and Q-like bands, increased
emission, and reduced Stokes shift, similar to PF32h. The
absorption spectrum of N-PF32h appears at more red-shifted
wavelength region, but distinct Q-like bands and fluorescence are
not observed; its 1H NMR spectrum does not exhibit a newly
formed aromatic current flow due to its structurally fixed figure-
eight conformation.
The excited-state dynamics of neutral PF32h, DCP32h, and
N-PF32h, as well as their deprotonated species, were examined
(Figure 49). Femtosecond TA spectroscopy shows that the S1-
state lifetimes of neutral and deprotonated PF32h are about 10
and 300 ps, respectively. An almost 30 times increased excited
state lifetime of deprotonated PF32h and its intensified
fluorescence support that the nonradiative decay rate in the S1
state of anionic PF32h is significantly reduced with regard to the
radiative rate constant. The S1-state lifetimes of neutral and
anionic DCP32h are obtained as 108 and 182 ps, respectively.
Similar to these photophysical features obtained from the 1H
NMR and steady-state experiments, the long-lived excited-state
lifetime of anionic DCP32h is ascribed to the increased degree of
aromaticity as compared with its neutral species. For N-PF32h,
the S1-state lifetime increases from 16 (neutral species) to 19 ps
(deprotonated species). Although TBAF can eliminate the N−H
protons of N-PF32h, the fixed molecular conformation possibly
does not allow huge structural changes such as unfolding. The
increased S1-state lifetimes of anionic PF32h and DCP32h as
compared with those of their neutral species can be considered as
evidence for conformational modifications toward an increased
degree of aromaticity.
For quantitatively evaluating the aromaticity, NICS values52
are obtained at various positions within the molecular cavity of
Figure 48. Steady-state absorption (solid line) and fluorescence (dotted [32]heptaphyrins (Figure 50). The calculations also support the
line) spectra of neutral (black line) and deprotonated (red line) species increased aromaticity of deprotonated PF32h and DCP32h by
of PF32h, DCP32h, and N-PF32h in CH2Cl2. Reprinted with
permission from ref 69. Copyright 2014 Royal Society of Chemistry.
elimination of NH protons. The NICS value estimated at the
center of the macrocycle (−7.25 ppm) of the π-conjugation
pathway of anionic PF32h supports its strong aromaticity, while
reality, from its crystal obtained from a pentane solution, PF32h that of neutral PF32h is obtained to be −1.79 ppm (Table 8).
is assumed to be a figure-eight structure.66 With the addition of The NICS values for the neutral and deprotonated states of
excess base, the absorption spectrum of PF32h changes, DCP32h are estimated to be around −7 ppm, which are in
featuring an intense, sharp B-like band and Q-like bands in the agreement with the increased aromaticity caused by the change
NIR region, and high fluorescence quantum yield is observed. in conformations by elimination of N−H protons. On the other
The small Stokes shift, <200 cm−1, suggests the comformational hand, the NICS values for the neutral and deprotonated states of
rigidity of deprotonated PF32h in the S1 state, and these spectral N-PF32h are calculated as 0.29 and −0.56 ppm, respectively.
features are typical characters of aromatic expanded porphyrins; Also, the HOMA values41 for the neutral and deprotonated
hence, anionic PF32h is suggested to exhibit a Möbius aromatic states of PF32h are estimated as 0.58 and 0.74, respectively. The
topology based on the previous studies.37,62,63,66,68,139−141 large increase in the HOMA value by elimination of NH protons
The elimination of N−H protons leads to the interruption of indicates newly formed π-conjugation and enhanced aromaticity.
intramolecular hydrogen bonds; hence, the figure-eight topology By deprotonation, as displayed in Table 8, the TPA value of
changes to the unfolded Möbius structure. The 1H NMR PF32h considerably increases, while that of DCP32h slightly
spectrum of DCP32h in CDCl3 is broad at 293 K but becomes increases; these results support that the deprotonation of
narrow, thereby permitting the assignment of its Möbius [32]heptaphyrins results in enhanced aromaticity. By elimi-
topologies at low temperature.66 In CH2Cl2, intense and sharp nation of NH protons with base, nonaromatic PF32h with a
B- and Q-like bands are observed in the absorption spectrum of highly distorted molecular structure and nonaromatic character
DCP32h, and fluorescence is observed at the NIR region; these undergoes structural modifications, affording unfolded Möbius
spectral properties describe that the major conformer existent in aromatic topology, and Möbius aromatic DCP32h exhibits
CH2Cl2 is a Möbius aromatic species. In line with this conjecture, stronger aromaticity caused by structural fixation. These features
its X-ray structure exhibits a Möbius topology.66 Deprotonation are accompanied by intense and red-shifted B-like bands, well-
of DCP32h results in the enhancement of degree of aromaticity, resolved Q-like bands, long-lived excited states, and increased
as revealed by 1H NMR, absorption, and emission spectra. The TPA values. On the other hand, these changes are not observed
2288 DOI: 10.1021/acs.chemrev.6b00313
Chem. Rev. 2017, 117, 2257−2312
Chemical Reviews Review

Figure 49. TA spectra and decay profiles of neutral (left) and deprotonated species (right): PF32h (a, b), DCP32h (c, d), and N-PF32h (e, f).
Reprinted with permission from ref 69. Copyright 2014 Royal Society of Chemistry.

for N-PF32h. Accordingly, deprotonation is demonstrated to be 51 reveals the steady-state absorption spectrum of PF36O in
a new method for dramatically controlling the topology of CH2Cl2; two broad bands are observed without distinct Q-like
[32]heptaphyrins and, hence, their aromatic character. bands, which are typical photophysical properties of nonaromatic
3.2.3. Deprotonation Effects in Larger Expanded molecular systems. With the addition of TBAF (40 equiv), the
Porphyrins. As their molecular size is not very large (<40π absorption spectrum of PF36O changes, exhibiting intense and
electrons), [4n] and [4n+2]π neutral porphyrinoids typically sharp B- and Q-like bands, and fluorescence appears (Figure 51).
exhibit antiaromatic and aromatic chemical characteristics, These photophysical features are characteristic of typical
respectively, according to the [4n+2] Hückel rule. Nevertheless, aromatic expanded porphyrins. Hence, the newly formed species
with increasing molecular size, the molecular framework is so are assumed to be a monoanionic octaphyrin with a Möbius
flexible that the overall structures are distorted so as to exhibit topology, PF36O−.
nonaromatic features, regardless of the number of π electrons. The obviously distinguished β-CH protons of PF36O−, eight
For retrieving Hückel or Möbius aromaticity in nonaromatic outer, four boundary, and four inner β-CH protons in the 1H
expanded porphyrins containing [4n] or [4n+2]π electrons, NMR spectrum (Figure 52), indicate the newly generated
various methods have been proposed, including protonation and aromatic ring current for the anionic species. Hence, it is
deprotonation. In this section, the deprotonation-induced suggested that the half-twisted Mö bius aromatic species,
conformational changes of PF36O with TBAF for affording PF36O−, is formed by considering the number of π electrons
the Möbius aromatic and Hückel antiaromatic compounds were (36). This assignment is in agreement with C1 symmetric spectral
demonstrated, depending on the amount of TBAF addition pattern.
(Chart 7).142 With addition of excess TBAF, further spectral changes are
PF36O is well-known to be a nonaromatic expanded observed. Intense absorption bands and a very weak, broad
porphyrin, with a highly distorted molecular structure.108 Figure absorption tailing are observed at the expense of the absorption
2289 DOI: 10.1021/acs.chemrev.6b00313
Chem. Rev. 2017, 117, 2257−2312
Chemical Reviews Review

Figure 50. Optimized molecular structures of neutral and deprotonated PF32h (a, b), DCP32h (c, d), and N-PF32h (e, f). All the meso-substituents and
pyrrolic β-protons are omitted for clarity. Reprinted with permission from ref 69. Copyright 2014 Royal Society of Chemistry.

Table 8. NICS, HOMA, Excited Singlet-State Lifetimes, and Chart 7. Deprotonation of PF36O; Monoanionic PF36O−
TPA Cross Section σ(2) of PF32h, DCP32h, and N-PF32h69 and Dianionic PF36O2− (Reprinted with Permission from Ref
142; Copyright 2016 Royal Society of Chemistry)
[32]heptaphyrin NICS (ppm) HOMA τS (ps) σ(2) (GM)
PF32h −1.79 0.58 9.5 1300
PF32h-TBAF −7.25 0.74 290 2600
DCP32h −6.71 0.68 108 2500
DCP32h-TBAF −7.44 0.72 182 2800
N-PF32h +0.29 0.60 16
N-PF32h-TBAF −0.56 0.63 19

bands due to PF36O−, and emission is not observed (Figure


51).142 Solid-state analysis showed that further deprotonated
PF36O2− assumes a planar square topology (Figure 53). Two
tetrabutylammonium cations are located at the negatively
charged nitrogen of PF36O2− for charge compensation. Thus,
PF36O2− could be assigned as dianionic [36]octaphyrin. This
topological analysis is in agreement with that of PF36O− as a
monoanionic species. This planar conjugation pathway with 36π
electrons implies that PF36O2− should be a Hückel antiaromatic
species. The absorption and fluorescence spectra of PF36O2− are
consistent with the above analysis. This planar square topology of
PF36O 2− is reminiscent of antiaromatic cyclooctafur-
an(1.1.1.1.1.1.1.1) reported by Anand and co-workers.143 The its molecular symmetry. The spectroscopic recovery from the
planar topology of PF36O2− is formed by Coulombic repulsions Möbius aromatic PF36O− and Hückel antiaromatic PF36O2− to
between the negatively charged nitrogens in the molecular cavity. nonaromatic PF36O by titration with TFA is observed (Figure
The intensified B-like bands of PF36O2− could be explained by 51).
2290 DOI: 10.1021/acs.chemrev.6b00313
Chem. Rev. 2017, 117, 2257−2312
Chemical Reviews Review

minimized molecular structures of PF36O, PF36O−, and


PF36O2−. The HOMA values of PF36O, PF36O−, and
PF36O2− are calculated as 0.56, 0.77, and 0.52, respectively.
The higher HOMA value for PF36O− suggests a smoothly
twisted aromatic π-conjugation pathway, while the small HOMA
value of PF36O2− with planar square molecular topologies
indicates a [4n] Hückel antiaromatic character.
In the MO diagram, four almost-degenerated FMOs for
PF36O− are calculated, which are characteristic of aromatic
expanded porphyrins, and the energy gap between HOMO and
LUMO is calculated to be ∼1.3 eV. On the other hand, the
calculated HOMO−LUMO gap of PF36O2− is relatively small,
0.95 eV,142 characteristic of antiaromatic porphyrinoids.
In addition, the excited-state dynamics of PF36O, PF36O−,
and PF36O2− are explored by femtosecond TA measurements
(Figure 56). TA spectroscopic measurements reveal the excited-
state lifetimes of 8.6, 50, and 9.8 ps for PF36O, PF36O−, and
PF36O2−, respectively. The >5 times longer S1-state lifetime of
PF36O− and its enhanced fluorescence suggest the aromaticity
of PF36O−. On the other hand, the excited-state dynamics of
antiaromatic PF36O2− show two decay components with
ultrafast and relatively slow decay components (0.4 and 9.8
ps). As described in the Introduction, the short time component
for PF36O2− is attributed to the relaxation to an optically dark
state, indicated by the calculated excitation energies around 2600
nm. The relatively long time constant possibly corresponds to
Figure 51. Absorption spectral changes during spectroscopic titration of the internal conversion from the optically dark state to the
PF36O with TBAF (a) from 0 to 40 and (b) from 40 to 7000 equiv in ground state. On the basis of our previous studies, the TA spectra
CH2Cl2. Inset is the spectroscopic recovery by TFA from fully for aromatic expanded porphyrins typically exhibit intense GSB
deprotonated species PF36O2− to neutral PF36O. Reprinted with
signals with relatively weak ESA signals. These aromatic signals
permission from ref 142. Copyright 2016 Royal Society of Chemistry.
are observed in the TA spectra of PF36O−, while those of
PF36O2− reveal opposite phenomena. These photophysical
The calculated NICS values52 of the Möbius aromatic species
features of PF36O2− support its antiaromaticity.
are in good agreement with experimental data if magnetic fields
In conclusion, by elimination of N−H protons with TBAF,
are orthogonally applied with regard to the main π-conjugation
highly distorted nonaromatic PF36O with a figure-eight
pathway in twisted conformation. The negative NICS(0) value of
topology undergoes structural changes to the half-twisted
PF36O− at the central position of the cavity is estimated, while
Möbius aromatic species PF36O− and square Hückel anti-
that of PF36O2− is calculated as a positive value, suggesting its
aromatic form of PF36O2−, depending on the amount of base.
antiaromaticity. For visualizing delocalized π electrons, the ACID
PF36O2− is revealed to assume a planar square molecular
plots137 of PF36O− and PF36O2− are also calculated.142 The π-
topology as a new conformation for meso-aryl [36]octaphyrin.
conjugated electronic circuit of PF36O− has definite electron
density in the smoothly twisted Möbius topology; a definite,
clockwise current (Figure 54) supports its aromaticity, while a 4. CONTROL OF AROMATICITY BY VARIOUS
counterclockwise current of PF36O2− indicates its antiaromatic SOLVENTS
nature (Figure 55). Moreover, the HOMA analysis by using the The molecular structural control of expanded porphyrins still
main π-conjugation pathways is performed from the energy- remains important not only for changing their aromatic character

Figure 52. 1H NMR spectrum of PF36O− in CDCl3 at 25 °C. Peaks marked with * are due to residual solvents, tetrabutylammonium ion, and
impurities. Reprinted with permission from ref 142. Copyright 2016 Royal Society of Chemistry.

2291 DOI: 10.1021/acs.chemrev.6b00313


Chem. Rev. 2017, 117, 2257−2312
Chemical Reviews Review

Figure 53. X-ray crystal structure of PF36O2−. (a) Top view and (b) side view. The thermal ellipsoids are represented at 50% probability. Solvent
molecules and counter cations are omitted for clarity. Reprinted with permission from ref 142. Copyright 2016 Royal Society of Chemistry.

Figure 54. (a) NICS values of Bq ghost atoms for PF36O−. (b) ACID
plot of PF36O− at an isosurface value of 0.05. The external magnetic
field is applied orthogonal to the molecule plane with vector pointed Figure 55. (a) NICS values of Bq ghost atoms for PF36O2−. (b) ACID
toward the viewer. Reprinted with permission from ref 142. Copyright plot of PF36O2− at an isosurface value of 0.05. The external magnetic
2016 Royal Society of Chemistry. field is applied orthogonal to the molecule plane with vector pointed
toward the viewer. Reprinted with permission from ref 142. Copyright
but also for controlling their physical and chemical properties. In 2016 Royal Society of Chemistry.
the following sections, the solvent-dependent conformational
changes of expanded porphyrins are mainly investigated by
Dramatic solvent-dependent spectral changes are shown in the
spectroscopic methods in various solvents, which are related to
absorption measurements of PF32h (Figure 57). In the
the changes in the degree of (anti)aromaticity and electronic
absorption spectrum of PF32h in toluene, only a weak and
properties.
broad B-like band is observed and Q-like bands are totally absent.
4.1. Control of Aromaticity in Heptaphyrin by Solvent In ethyl ether, a broad and split B-like band appears in the visible
Here, the solvent polarity-dependent optical properties of PF32h region with extremely weak Q-like bands in the NIR region.
with respect to their conformations and aromaticity are described Especially, narrow and strong B- and Q-like bands gradually
(Chart 6).66 Importantly, the TA spectroscopy and global appear in THF. The absorption spectra are further changed in
analysis are employed for resolving complicated conformational highly polar solvent, DMSO, where both B- and Q-like bands
dynamics in solution via the detection of characteristic become red-shifted and their oscillator strengths are intensified.
photophysical features of nonaromatic and (anti)aromatic This tendency of spectral changes implies that the molecular
expanded porphyrins with their conformations. structure of PF32h is very sensitive to the surrounding medium.
2292 DOI: 10.1021/acs.chemrev.6b00313
Chem. Rev. 2017, 117, 2257−2312
Chemical Reviews Review

Figure 56. Femtosecond TA spectra and decay profiles of (a) PF36O, (b) PF36O−, and (c) PF36O2− in CH2Cl2. Reprinted with permission from ref
142. Copyright 2016 Royal Society of Chemistry.

the spectral shapes in THF. The absorption spectrum in ethyl


ether can be reproduced via the addition of toluene and THF in a
ratio of 6:1 (Figure 58); the huge difference between the two

Figure 57. Steady-state absorption and fluorescence (inset) spectra of


PF32h in various solvents measured at room temperature (296 K).
Reprinted with permission from ref 65. Copyright 2011 Wiley-VCH
Verlag GmbH & Co. KGaA.

On the basis of related studies, the steady-state absorption Figure 58. Reproduced (dotted line) and observed (solid line)
spectrum in highly polar solvent is affiliated to a Möbius aromatic absorption spectra of PF32h in ethyl ether. Observed absorption
conformer and that in less-polar solvent belongs to an spectra in toluene and THF were represented at the molar ratio of 6:1.
antiaromatic species.38,66,94 Hence, an increase of solvent polarity Reprinted with permission from ref 65. Copyright 2011 Wiley-VCH
Verlag GmbH & Co. KGaA.
is suggested to trigger the structural changes of PF32h from
antiaromatic to aromatic topologies.
The fluorescence is only observed in polar solvents such as spectra is a small divergence of spectral shift, owing to differences
THF and DMSO (Figure 57). The overall shapes of fluorescence in solvation. Hence, two types of molecular topologies coexist in
spectra in THF and DMSO are quite similar, suggesting that the ethyl ether.
emitting species are quite similar in both solvents. However, the Femtosecond TA measurements depending on the solvent
intensity of vibronic sideband is stronger in more-polar DMSO, polarity and excitation wavelength provide in-depth information
which suggests that the conformer in DMSO is strongly fixed so for conformational dynamics of PF32h. The TA spectra in
that the structural changes from the initial Franck−Condon to toluene exhibit broad GSB and intense ESA bands (Figure 59a).
the relaxed excited-state geometry are relatively small. Within 3 ps, the ESA band becomes sharp and blue-shifted. This
A closer look at the newly generated signals in ethyl ether, split kind of spectral change arises from vibrational relaxation
B-like bands and weak Q-like bands, reveals that they resemble including both intra- and intermolecular processes due to
2293 DOI: 10.1021/acs.chemrev.6b00313
Chem. Rev. 2017, 117, 2257−2312
Chemical Reviews Review

Figure 59. (a) Femtosecond TA spectra, (b) decay-associated spectra, Figure 60. (a) Femtosecond TA spectra, (b) decay-associated spectra,
and (c) temporal profiles of PF32h in toluene measured at room and (c) temporal profiles of PF32h in THF measured at room
temperature (296 K). Excitation wavelength is 600 nm. Reprinted with temperature (296 K). Excitation wavelength is 650 nm. Reprinted with
permission from ref 65. Copyright 2011 Wiley-VCH Verlag GmbH & permission from ref 65. Copyright 2011 Wiley-VCH Verlag GmbH &
Co. KGaA. Co. KGaA.

constants of 10 and 60 ps correspond to the S1-state lifetimes in


photoexcitation at higher lying excited states.144,145 By global toluene and THF, respectively, this result gives direct evidence of
analysis, as shown in Figure 59b, two DAS are extracted. The the suggestion that two conformers are present in ethyl ether. At
DAS of ∼0.4 ps shows red-shifted and broad ESA, whereas that of a photoexcitation wavelength of 590 nm, interestingly, the TA
∼5 ps reveals blue-shifted ESA. These two DAS therefore spectra show spectral behaviors similar to the TA spectra by
represent the TA spectra before and after vibrational relaxation photoexcitation at 640 nm in ethyl ether (Figure 61d). There is
processes, respectively, not those of two different species. In only marginal difference: the excited-state population of the
other words, the time component of 0.4 ps can be regarded as the conformer in THF is a very small quantity. Hence, GSB is not
vibrational relaxation and the lowest excited-state lifetime is strong and the S1-state lifetime corresponding to that of the
estimated to be ∼5 ps in toluene. This fast S1-state lifetime is in conformer in THF is not shown in decay profiles, although three
line with the nonfluorescent photophysical feature in less-polar types of DAS are extracted from global analysis (Figure 61e). In
solvents. fact, the ground-state population of the conformer in THF is
The TA spectra measured in THF exhibit intense GSB but quite small (14%) in ethyl ether, as shown in the calculated
relatively weak ESA (Figure 60). In addition, with increasing vertical transitions (Figure 58). Finally, intense and sharp GSB at
delay time, spectral changes are not revealed. In global analysis, ∼685 nm and weak ESA are observed in the TA spectra of PF32h
only one DAS with the time constant of 65 ps is extracted, in DMSO (Figure 62), exhibiting incomplete excited-state
indicating that there is only one conformer in THF (Figure 60b). dynamics without spectral shift as observed in THF.
More complicated features appear in the TA spectra obtained in In addition, only one DAS is obtained from global analysis with
ethyl ether (Figure 61). By photoexcitation at 640 nm, at the a time component of 234 ps, which can be assigned to be the S1-
initial time delay, broad and sharp GSB and intense ESA are state lifetime in highly polar DMSO; the relatively long lifetime is
observed. With increasing time delay, these signals rapidly comparable to that of vinylene-bridged [26]hexaphyrin.54
decrease, whereas GSB at 640 nm slowly decays. Meanwhile, Generally, expanded porphyrins are adversely affected by
ESA at ∼610 nm and weak GSB of 690 nm gradually rise (Figure structural distortion to twisted form and exhibit short excited-
61a). These complex excited-state dynamics in ethyl ether verify state lifetimes. In this regard, PF32h appears to maintain a quite
that there are two different species, which are analogous to fixed molecular structure in highly polar solvents. This conjecture
conformers in toluene and THF. might be due to hydrogen bonding of DMSO.66,68,141
In detail, as the DAS with time constants of 0.5 and 10 ps Temperature-dependent absorption measurements support
exhibit similar spectral features to those observed in toluene the existence of two conformers in ethyl ether (Figure 63). With
(Figure 59), it can be said that these two spectra are attributed to a gradual decrease in temperature from 290 to 180 K, the split B-
the conformer in toluene. The DAS corresponding to the time like bands in the steady-state absorption spectra of the porphyrin
constant of 60 ps also shows a similar spectral feature to that in in ethyl ether increase, in addition to the growth of the shoulder
THF (Figure 60). Therefore, considering that the two time peak and bands in the NIR region. With decreasing temperature,
2294 DOI: 10.1021/acs.chemrev.6b00313
Chem. Rev. 2017, 117, 2257−2312
Chemical Reviews Review

Figure 61. (a, d) Femtosecond TA spectra, (b, e) decay-associated spectra, and (c, f) temporal profiles of PF32h in ethyl ether with photoexcitation at
(a−c) 640 nm and (d−f) 590 nm measured at room temperature (296 K). Reprinted with permission from ref 65. Copyright 2011 Wiley-VCH Verlag
GmbH & Co. KGaA.

Figure 63. Temperature-dependent absorption spectra of PF32h in


ethyl ether. Reprinted with permission from ref 65. Copyright 2011
Wiley-VCH Verlag GmbH & Co. KGaA.

on the other hand, the intensity of the split B-like bands at 600
nm decreases. With decreasing temperature, the increased band
is attributed to the absorption bands in THF, while the
diminished band corresponds to the B-like band in toluene.
The absorption spectrum (at 183 K) of expanded porphyrin in
ethyl ether shows similar optical features and peak positions to
those in THF. The interpretation of these temperature-
dependent spectroscopic behaviors is the thermodynamic
Figure 62. (a) Femtosecond TA spectra, (b) decay-associated spectra, equilibrium between conformers with decreasing temperature.
and (c) temporal profiles of PF32h in DMSO measured at room Quantum mechanical calculations qualitatively analyze the
temperature (296 K). Excitation wavelength is 675 nm. Reprinted with possible conformations and their aromaticity for PF32h (Figure
permission from ref 65. Copyright 2011 Wiley-VCH Verlag GmbH & 64). First, the conformers in toluene and THF are assumed to
Co. KGaA. have antiaromatic and aromatic topologies, respectively, based
2295 DOI: 10.1021/acs.chemrev.6b00313
Chem. Rev. 2017, 117, 2257−2312
Chemical Reviews Review

to 193 K,65 suggesting that figure-eight as well as double-sided


topologies appear to be maintained at 193 K. On the other hand,
the 1H NMR spectrum in THF-d8 reveals aromatic character
without temperature dependency.66 Hence, a less-symmetric and
highly distorted Möbius conformation exists in THF. In ethyl
ether-d10, temperature-dependent NMR spectra show significant
changes. The NMR spectrum at <250 K is similar to that
obtained in THF-d8, while the NMR spectrum is almost the same
as that in toluene-d8 at 293 K. These results are in accordance
with the temperature-dependent, steady-state spectra, indicating
that two conformers exist in ethyl ether. In other words, their
thermodynamic equilibrium is sensitive to temperature changes.
On the basis of the above results, the two coexisting
conformational isomers would be aromatic and antiaromatic
species in ethyl ether, which correspond to the conformers in
toluene and THF, respectively, as suggested by steady-state and
time-resolved spectroscopy. Typically, because aromatic species
Figure 64. ACID plots of (a) figure-of-eight-like and (b) Möbius are thermodynamically more stable than antiaromatic mole-
structures of PF32h. Current density vectors indicate counterclockwise cules,48,146 the equilibrium shift to the conformer in THF at
(top) and clockwise (bottom) overall currents, respectively, when we lower temperatures implies that it is more stable than the
look toward the opposite direction to the magnetic field vector conformer in toluene. In the previous study about PF28H, the
(clockwise currents are diatropic, while counterclockwise currents are aromatic topology became predominant at low temperatures in
paratropic). The magnetic field vector is applied as a large arrow in the the equilibrium dynamics between Möbius aromatic and Hückel
small figures at the right-hand side. Reprinted with permission from ref
65. Copyright 2011 Wiley-VCH Verlag GmbH & Co. KGaA.
antiaromatic conformers.62,63
The TPA values in several solvents reflect the changes in the
degree of aromaticity. Besides, as mentioned in the Introduction,
on their 1H NMR spectra. In a previously reported result, PF32h given that the two expanded porphyrins are comparable to each
has been reported to assume a figure-eight antiaromatic and a other under the same conditions, as compared to the
Möbius aromatic (DCP32h) geometry, based on the solid-state antiaromatic expanded porphyrin, the aromatic one exhibits
analysis;66 hence, they are utilized as initial structures for higher TPA cross section values.94 Hence, the TPA cross section
optimization. Even after optimization, the geometries are values have been frequently used in several research studies for
maintained.65 Then, vertical excitation transitions are calculated. investigating the degree of aromaticity and π electron
The calculated excitation energies for both aromatic and delocalization in expanded porphyrins.94 The TPA cross section
antiaromatic topologies are in good agreement with the steady- values of PF32h increase from 2000 to 4700 GM as solvents are
state absorption spectra of expanded porphyrins.65 Hence, it is changed from toluene to DMSO. The increased TPA cross
quite relevant to consider these aromatic and antiaromatic section values in THF result in the Möbius topology. Thus, in
conformers as appropriate molecular structures in both solvents. ethyl ether, a relatively small increase in the TPA cross section
Here, notably, the vertical excitation transition of zero oscillator value is due to a small portion of the conformer in THF. The
strength in the NIR region, an optically dark state, appears from huge TPA value in DMSO indicates the predominance of the
the conformer in toluene. On the basis of the previous studies, aromatic conformer in DMSO, which is structurally fixed, caused
the existence of an optically dark state is considered to be a by hydrogen bonds between solvents and solutes.
unique photophysical property of antiaromatic expanded The transient absorption spectra of the conformers in toluene
porphyrins, which play a significant role in energy-relaxation and THF exhibit spectral features of typical aromatic and
processes.54 The shapes of MOs as well as energy levels for both antiaromatic expanded porphyrins, respectively. On the basis of
aromatic and antiaromatic topologies reveal a similar the previous studies of various expanded porphyrins, the
trend.54,62,63 photophysical properties have been suggested to be related to
The ACID137 and NICS52 calculations illustrate the degree of their aromatic character in terms of the absorption and
aromaticity in these forms quantitatively. While the direction of fluorescence spectra, excited-state dynamics, and TPA cross
current density of the antiaromatic species has counterclockwise section values.94
currents, the aromatic species exhibit clockwise currents. Moreover, a certain relationship has been established between
Notably, in ACID calculations, the same molecule exhibits the shape of the TA spectrum and the aromaticity.54
opposite directions depending on the molecular topologies even Antiaromatic expanded porphyrins typically reveal broad and
though conformational differences are small. In addition, the huge ESA relative to the GSB, as observed for the conformer in
NICS(0) values are estimated to be positive and negative, toluene, while aromatic species exhibit strong and sharp GSB as
supporting their antiaromatic and aromatic character, respec- compared to ESA, as can be observed in the conformer in THF.
tively. These results propose that a small change in conformation This trend can be observed in the case of vinylene-bridged [26]-
results in a significant change in (anti)aromaticity. The variation and [28]hexaphyrins.54 Nevertheless, at this stage, the reasons as
of solvent polarity is sufficient for triggering such a small to why the transient absorption spectra appear to depend on
structural modification. In accordance with the ACID plots and (anti)aromaticity are not clear. As demonstrated by TA
NICS values, similar aromaticity is observed in the 1H NMR spectroscopy, broad ESA from the conformer in toluene exhibits
spectra of PF32h. The 1H NMR spectrum at room temperature a spectral shift to the short-wavelength region at an early time
in toluene-d8 reveals a paratropic ring current.66 In addition, no domain; these observations are characteristic of vibrational
significant change is observed with decreasing temperature down relaxation. When excited-state electronic populations are formed
2296 DOI: 10.1021/acs.chemrev.6b00313
Chem. Rev. 2017, 117, 2257−2312
Chemical Reviews Review

in high vibrational states, additional absorption bands can be spectra, the huge TPA value, and a relatively long excited-state
intuitively expected in the low-energy region, for instance, hot lifetime.
band absorption according to the Franck−Condon principle, In conclusion, PF32h shows solvent dependency with regard
which predicts the absorption or fluorescence spectra by to various photophysical properties including absorption,
vibrational overlap between the initial and final states.100 Thus, fluorescence, 1H NMR, TA, and TPA measurements. All the
the ground-state absorption spectrum becomes broader in this photophysical results demonstrate that this molecule is switched
situation; hence, the feature of the transient absorption spectra in from antiaromatic to aromatic with increasing solvent polarity
toluene is assigned as an indicator for vibrational relaxation. On from toluene to DMSO. On the basis of computational
the other hand, in the TA in THF, there is no mark of vibrational calculations, this feature is ascribed to structural changes between
relaxation processes. It is assumed that this difference arises from Möbius aromatic and Hückel antiaromatic topologies. These two
the particular electronic state of the conformer in toluene. conformers coexist in ethyl ether, and the absorption and
The antiaromatic conformer in toluene, unlike the conformer transient absorption spectra are observed as a mixture of the
in THF, shows one-photon forbidden transition at ∼2250 nm in conformers found in THF and toluene, respectively. The excited-
the calculated vertical transitions, which is due to the existence of state lifetimes of the two conformers can be separated by
the optically dark state. Although the existence of the optically selective photoexcitation. In the TA spectra in toluene,
dark state cannot be probed by the ground-state absorption vibrational relaxation process becomes manifest, which repre-
spectra, ultrafast energy relaxation dynamics in toluene reveal the sents broad ESA at early time delays, while such phenomena
presence of the optically dark state. The decay profiles in THF cannot be observed in polar solvents.
exhibit a decay component (65 ps) with photoexcitation at the B
state (SB). This relatively long time constant is originated from
5. REVERSAL OF AROMATICITY IN THE EXCITED
the Q state (SQ) after internal conversion processes from SB to SQ STATE
states (<200 fs). In contrast, the decay profiles of the conformer During the last several decades, various attempts have been made
in toluene show two time components of 0.4 and 5 ps with for quantifying the magnitude of aromaticity in π-conjugated
photoexcitation at SB transition. Because the internal conversion cyclic compounds because their chemical properties and
processes from the SB to SQ states would be very fast in THF, an reactivities are predominantly governed presumably by aroma-
extra decay time component of <1 ps in toluene indicates the ticity or antiaromaticity. In addition to completely planar
presence of a low-lying state below the SQ state. The state with a molecular systems, various multidimensional aromatic systems
time component of 5 ps is substantially shorter than that of 65 ps have been widely synthesized. Although a multitude of
in THF, suggesting that the energy relaxation to the ground state experimental and quantum chemical parameters have been
of the conformer in toluene becomes more rapid than that of the employed for the evaluation of aromaticity in annulenes,
conformer in THF owing to the energy-gap law.95,116 In fact, photophysical properties, such as absorption and fluorescence
comparative excited-state dynamics have already been revealed behaviors, excited-state lifetimes, electronic structures, and NLO
from typical aromatic and antiaromatic congeneric porphyr- properties have been found to be reliable experimental indices for
inoids.54 As described in the Introduction, the aromatic evaluating the degree of aromaticity in a comparable set of [4n]
expanded porphyrins show relatively long decay components, and [4n+2]π expanded porphyrins, as described in a previous
section. Nevertheless, most of the photophysical studies on
while antiaromatic congeners reveal two decay components of
molecular aromaticity have been limited to ground electronic
<1 ps and ∼10 ps. In addition, the one-photon forbidden
states. Although excited-state aromaticity is one of the most
transitions are calculated from antiaromatic porphyrinoids; these
interesting concepts for photostability, chemical reactivity, and
forbidden transitions are observed as weak and broad absorption
applications in photonic devices, it has only rarely been
tailings at NIR region. Thus, the energy-relaxation dynamics of
investigated. In 1972, Colin Baird reported the aromaticity rule
antiaromatic [32]heptaphyrin can be explained by similar for the lowest triplet state by the perturbation molecular orbital
processes as those mentioned above. The time constants of <1 theory.46 Baird reported that ground-state Hückel’s aromaticity
ps from antiaromatic molecules are estimated to be internal is reversed in the lowest triplet state: annulene with [4n]π
conversion processes from the SB to the optically dark state, electrons exhibits aromatic character, whereas those with [4n
judging from the fact that a short component is nearly the same +2]π electrons exhibit antiaromatic character, in contrast to that
for both typical antiaromatic porphyrinoids and the toluene observed in their ground states.
conformer of PF32h. Since the report of reversed Hückel’s aromaticity in the lowest
The characteristic photophysical features for antiaromatic triplet state by Baird, numerous theoretical studies related to
expanded porphyrins, such as intense and broad ESA, are caused Baird’s rule have been reported.46−48,51,147−158 Although Baird’s
by the accumulation of excess vibrational energy below the SQ pioneering achievement has only been confirmed by extensive
state. This optically dark state is caused by the characteristic MO theoretical studies, such as NICS calculations, the reversal of
energy and symmetry; that is, the aspect of electronic structure aromaticity in the lowest triplet state has not yet been
that governs the spectroscopic features of antiaromatic species; experimentally demonstrated. Thus, this section aims to address
thus, antiaromatic species can be differentiated from aromatic the spectroscopic signatures for supporting the reversal of
ones.39,54,63 In addition, the TA features possessing enough aromaticity in the lowest triplet state of [4n] or [4n+2]π
vibrational energy within the optically dark state can be an expanded porphyrins, and their characterization would contrib-
experimental indicator for antiaromatic species. On the basis of ute to the development of synthetic protocols, where aromatic
our concept, the TA spectral features of PF32h in highly polar stability and reactivity possibly play critical roles.
solvents show aromatic character. Although molecular structures Generally, it is difficult to synthesize a comparable set of stable
are not obtained but only estimated in DMSO, another Möbius [4n]π and [4n+2]π congeners. Recently, however, neutral and
conformation with enhanced rigidity is expected via hydrogen stable [4n]/[4n+2]π porphyrinoids have been reported.
bonds between solvents and solutes based on the 1H NMR Compared to other π-electronic systems, expanded porphyrins
2297 DOI: 10.1021/acs.chemrev.6b00313
Chem. Rev. 2017, 117, 2257−2312
Chemical Reviews Review

are advantageous in that they can be deprotonated or protonated Q-like bands, features characteristic of aromatic porphyrinoids,
on the electron oxidation or reduction of macrocycles. These are observed in the absorption spectrum of R26H in the S0 state.
molecular properties have enabled expanded porphyrins to be a In contrast, R28H affords relatively broad, weak absorption
good sample for evaluation of the aromaticity and examination of bands in the visible region with a smeared tail in the NIR region
Baird’s aromaticity. (Figure 65).
Expanded porphyrins are one of the most suitable molecular
entities for investigating the molecular aromaticity because of
their excellent photostability and optical properties, as well as an
easy synthesis of aromatic or antiaromatic congeners by simple
two-electron redox reactions. As described in the previous
section, as compared to aromatic expanded porphyrins, their
antiaromatic congeners exhibit (i) broad and weak absorption
spectra, (ii) free of emission, (iii) low TPA cross section values,
and (iv) an optically dark S1 state, which results in a rapid decay
to the ground state. Interestingly, the spectral shapes and
intensities in the TA spectra of aromatic and antiaromatic species
are contradictory to each other in a comparable set of [4n] and
[4n+2]π expanded porphyrins. Accordingly, in line with Baird’s
rule, the spectroscopic signatures can be expected to be reversed
in the excited states, which will provide experimental evidence on
the reversal of aromaticity in the excited states. This section
Figure 65. Absorption spectra of R26H (black) and R28H (red) in
covers the reversal of aromaticity in the lowest singlet and triplet toluene. Reprinted with permission from ref 71. Copyright 2015
excited states based on the TA measurements of various [4n] or Macmillan Publishers.
[4n+2]π Möbius- and Hückel-type expanded porphyrins.
5.1. Reversal of Aromaticity in the Lowest Triplet State As compared with free-base hexaphyrins,93 the S1-state
For probing the reversal of aromaticity in the T1 state as lifetimes of R26H and R28H are significantly reduced because
suggested by Baird, the T1-state aromaticity of a pair of of efficient intersystem crossing due to the heavy-atom effect of
bisrhodium [26]- and [28]hexaphyrins, R26H and R28H, the coordinated rhodiums in hexaphyrins. The femtosecond TA
respectively, is considered (Chart 8).159 Although these two spectra of R26H are fitted with double exponential functions
with time constants of 17 ps and ≥5 ns, which correspond to the
Chart 8. Molecular Structures of R26H (Left) and R28H S1- and T1-state lifetimes, respectively (Figure 66). By the global
(Right) (Reprinted with Permission from Ref 71; Copyright analysis of the whole TA spectra via SVD analysis, the DAS of
2015 Macmillan Publishers) two transient components assigned to the S1 and T1 states,
respectively, are obtained, as illustrated in Figure 67. Because the
DAS consists of only ESA and GSB (free of SE due to the heavy-
atom effect), the ESA spectrum of the T1 state could be extracted
from the GSB and DAS of the T1 state. Similarly, the two DAS of
R28H are attained, which are hypothesized to stem from the S1
and T1 states with time constants of 11 and 195 ps, respectively.
Because of free-of-fluorescence in R28H, the ESA spectrum of
the T1 state was estimated by subtracting the GSB from the DAS.
For quantitatively comparing differences of absorbance
between the S0 and T1 states, the extinction coefficient for
molecules have similar molecular framework, one compound has each state was estimated, as shown in Figure 68. From this
[4n+2]π electrons and the other has [4n]π electrons. The analysis, the ESA spectra of the T1 states of R26H and R28H
absorption spectral features and intensities in the ground and strongly indicate the reversal of aromaticity in the T1 state, so-
excited state of the Hückel aromatic (R26H) and antiaromatic called Baird’s rule. Particularly, the ESA spectrum of the T1 state
(R28H) congeners exhibit obvious distinctions. In this regard, in R26H exhibits relatively broadened and reduced absorption
R26H and R28H are the best testbeds for which the applicability spectral features, compared to the absorption spectrum of the S0
of Baird’s rule should be explored, for the following reasons: (i) state. This ESA spectrum is close to the absorption spectra of S0
coordinated rhodium metals that make the almost-planar state in typical antiaromatic porphyrinoids.38 This observation
structures rigid; (ii) chemically stable species; (iii) both [26]- proposes that the T1 state of R26H is presumably antiaromatic.
and [28]-electronic structures, which are [4n]π (Hückel Contrary to what was observed for a 26π electron species, the
antiaromatic) and [4n+2]π (Hückel aromatic) congeners, ESA spectrum of the T1 state in R28H exhibits relatively narrow
respectively; and (iv) coordinated rhodium metal that should and intensified spectra versus that of the S0 state; this ESA
promote the intersystem crossing processes required to generate spectrum resembles that of typical aromatic porphyrinoids.160
the corresponding T1 states due to heavy atom effect while (v) For rationalizing the contrasting feature in the absorption
not substantially affecting the spectroscopic properties due to the spectra, the vertical excitation energies of R26H and R28H were
closed-shell electronic structures of the rhodium(I) cen- calculated. The MO structures and calculated vertical energy
ters.71,159,160 transitions in the S0 and T1 states display a sharp contrast (Figure
R26H and R28H show Hückel aromatic and antiaromatic 69). As discussed in the Introduction,160 the absorption spectra
characters, respectively, in the S0 state, as is evident from 1H of the S0 state in aromatic R26H can be comprehended by
NMR and NICS results.159 Two strong B-like and several weak Gouterman’s four-orbital model.55 Conversely, the calculated
2298 DOI: 10.1021/acs.chemrev.6b00313
Chem. Rev. 2017, 117, 2257−2312
Chemical Reviews Review

Figure 68. Ground- (black lines) and triplet-state (red lines) absorption
spectra of (a) R26H and (b) R28H plotted as a function of extinction
coefficients. The red spectra are obtained by summation of the ground-
state absorption and scaled decay-associated TA spectra. The asterisk
(*) indicates experimental error peak due to different spectral
resolutions between the two spectrometers used in ground-state
absorption and TA measurements. Reprinted with permission from
ref 71. Copyright 2015 Macmillan Publishers.

such as HOMO to LUMO and HOMO−1 to LUMO+1


transitions. These calculated transitions in the S0 state are
completely switched in the T1 state. Hence, the calculated
vertical energy transitions of R26H in the T1 state are similar to
those of the S0 state of R28H. This contrasting feature in the
absorption spectra is also observed for R28H. Now, the T1 state
is similar to the S0 state of R26H. The complete reversal in
Figure 66. Femtosecond TA spectra of (a) R26H and (b) R28H in spectral features observed in the MO diagram as well as the
toluene with the photoexcitations at 590 and 505 nm, respectively.
estimated vertical transitions are completely in line with the
Reprinted with permission from ref 71. Copyright 2015 Macmillan
Publishers. aforementioned conclusion, namely, that the bisrhodium
hexaphyrins exhibit the reversal of Hückel (anti)aromaticity in
their T1 states.
transitions of R28H in the S0 state show the lowest-energy For demonstrating the proposed reversal of (anti)aromaticity
transition in the NIR region, which has a weak oscillator strength in the T1 state of bisrhodium hexaphyrins, the NICS52 values and
close to zero, with relatively small B-like transitions in the visible ACID137 plots were calculated both in the ground and T1 states
region, which is contributed to completely forbidden transition (Figure 70 and Table 9). The NICS(1) values were estimated

Figure 67. TA contour maps and ground-state absorption and decay-associated TA spectra of R26H and R28H. (a) TA contour map of R26H in
toluene. (b) TA contour map of R28H in toluene. (c) Ground-state absorption spectra of R26H in toluene shown as black line. The decay-associated
TA spectra of R26H in toluene are shown as red (17 ps) and blue (>5 ns) lines. (d) Ground-state absorption spectra of R28H in toluene shown as black
line. The decay-associated TA spectra of R28H in toluene are shown as red (11 ps) and blue (195 ps) lines. Reprinted with permission from ref 71.
Copyright 2015 Macmillan Publishers.

2299 DOI: 10.1021/acs.chemrev.6b00313


Chem. Rev. 2017, 117, 2257−2312
Chemical Reviews Review

Table 9. Calculated NICS(1) Values at Two Positions of


R26H and R28H in the Ground and Lowest Triplet States (All
Values Calculated Using the Geometry-Optimized Structures
at B3LYP/LANL2DZp Level Based on X-ray Crystallographic
Structures)71
R26H R28H
positions S0 T1 S0 T1
NICS(1) a −14.58 15.73 18.41 −14.00
b −10.85 15.50 18.10 −10.88

because the NICS(0) values could be perturbed by the sigma


bonds in the molecules.161 In the case of R26H, the NICS(1)
value at the center of the macrocycle is evaluated as a negative
value in the S0 state, which is in agreement with their aromatic
features in the experimental results. On the other hand, in the T1
state of R26H, a positive value is observed, which is in line with
switching from an aromatic to antiaromatic molecule by going
from S0 to T1 state (Table 9). Contrary to R26H, for R28H, a
positive NICS(1) value in the S0 state and a negative NICS(1)
value in the T1 state are estimated. Similarly, reversed sign of
NICS(1) values at the different positions of R26H and R28H in
the T1 state are estimated as compared to those in the S0 state.
Because the same sign and slight differences in the NICS values
are estimated at various positions, the NICS values are believed
to accurately represent the degree of the global aromaticity rather
than local aromaticity caused by the respective pyrrole rings.
The ACID plots are in agreement with the above analysis. The
overall direction of ring-current flow is switched from clockwise
(aromatic) to counterclockwise (antiaromatic) in the S0 and T1
Figure 69. Vertical excitation energy data calculated using TD-DFT states of R26H, respectively. Similar to R26H, in the case of
(blue) compared with the (a) ground- and (b) triplet-state absorption R28H, the direction of ring-current flow is also switched from
spectra (black) of R26H and R28H at B3LYP/LANL2DZP level. (The counterclockwise to clockwise (Figure 70 and Table 9).
asterisk (*) indicates experimental error peak due to different spectral Accordingly, these calculated results also show an agreement
resolutions between the two spectrometers used in ground-state with the spectroscopic observations.
absorption and TA measurements.) Reprinted with permission from Calculated HOMA values also support the suggested
ref 71. Copyright 2015 Macmillan Publishers. hypothesis related to the aromaticity in the T1 state.41,51,71 For
analyzing changes in geometry between the S0 and T1 states in

Figure 70. Isosurfaces (yellow) and current density vectors (small green arrows) calculated by ACID for R26H (left) and R28H (right) in the ground
(bottom) and lowest triplet state (top) at B3LYP/LANL2DZP level. Large arrows correspond to schematic representations for the overall ring-current
densities induced by external magnetic fields. Reprinted with permission from ref 71. Copyright 2015 Macmillan Publishers.

2300 DOI: 10.1021/acs.chemrev.6b00313


Chem. Rev. 2017, 117, 2257−2312
Chemical Reviews Review

more detail, the HOMA values for all possible 26π- and 28π- region, with an extremely weak absorption tail in the NIR region
conjugation pathways of R26H and R28H, respectively, were attributed to the optically dark state of P28H, is observed (Figure
thus evaluated. As compared with those of the S0 state, all HOMA 71).
values for the T1 state in R26H decrease, while those for the T1
state in R28H increase. Taken together, these results correspond
to the experimental results that the (anti)aromaticity of
congeneric bisrhodium hexaphyrins is switched in their T1
states. Surprisingly, nevertheless, the change in the absolute
HOMA value is smaller for R26H than for R28H because of the
instability of antiaromatic molecules, which can cause structural
distortion to nonaromatic forms.38,39,53,162
To sum up, the spectroscopic properties of bisrhodium
hexaphyrins in their respective T1 states, as compared with those
of the corresponding ground states, in combination with the
theoretical aromaticity indices, furnish experimental evidence on
the reversal of Hückel (anti)aromaticity. Furthermore, both
theoretical and experimental results are in accord with previous
studies, which proposed the reversal of (anti)aromaticity in the
T1 states of simple π-conjugated rings, i.e., cyclobutadiene,
benzene, and cyclooctatetraene.46−48,51,147−158
5.2. Reversal of Aromaticity in the Singlet Excited State
Since the report of aromaticity reversal in the lowest triplet state
proposed by Baird in 1972, numerous theoretical studies have
been conducted for extending Baird’s rule to the first singlet
excited state.47,48,152,163 In this context, in 2008, Karadakov
suggested that the aromaticity of annulene in the lowest singlet
excited state also could be reversed as compared with that in the Figure 71. Ground-state (black lines) and the lowest singlet excited-
ground state.47,48 Nevertheless, an experimental study for state absorption (red lines) spectra of (a) P26H and (b) P28H. The
demonstrating the reversal of (anti)aromaticity in the lowest asterisk (*) indicates instrumental error peak due to different spectral
resolutions of the spectrometers employed in absorption and TA.
singlet excited state by spectroscopic methods has been rarely
Reprinted with permission from ref 57. Copyright 2015 American
reported, although a number of studies have been dedicated Chemical Society.
toward the investigation of aromaticity in the excited state.
In this discussion, the lowest singlet excited-state aromaticity
of [26]- and [28]hexaphyrins is illustrated by the comparison of The femtosecond TA spectra of P26H exhibit single
their photophysical properties.57 Nevertheless, most of the meso- exponential decay with a time constant of 160 ps, corresponding
aryl [28]hexaphyrins have been reported to tend to be Möbius to the Qx state lifetime (Figure 72). By the global fitting of the
aromatic or figure-eight antiaromatic ones via structural
distortion in the solution phase.11,38,62,63 In contrast, internally
phenylene-bridged hexaphyrins, which are constrained to assume
spectacles-like conformation, can be a good candidate for
investigating singlet excited-state aromaticity (Chart 9).164 In

Chart 9. Molecular Structures of P26H and P28H (Reprinted


with Permission from Ref 57; Copyright 2015 American
Chemical Society)

this regard, phenylene-bridged [26]- and [28]hexaphyrins


(P26H and P28H, respectively) exhibiting Hückel aromaticity
and antiaromaticity in their ground states, respectively, were
utilized for 1H NMR and NICS calculations.71,72,164
Sharp and intense spectral features are observed in the ground- Figure 72. Ground-state absorption (black lines) and decay-associated
state absorption spectrum of P26H, similar to those exhibited by TA spectra (colored lines) of (a) P26H and (b) P28H in toluene.
aromatic porphyrins, except for large red-shifts. On the other Reprinted with permission from ref 57. Copyright 2015 American
hand, a relatively weak, broad absorption spectrum in the visible Chemical Society.

2301 DOI: 10.1021/acs.chemrev.6b00313


Chem. Rev. 2017, 117, 2257−2312
Chemical Reviews Review

whole TA spectra by SVD, the DAS of one transient species, Qx because of the limitation of quantum mechanical calculations of
states, can be evaluated, as shown in Figure 72. Because the DAS the excited state. Thus, the NICS values and ACID plots of
consists of GSB, SE, and ESA, the actual ESA spectrum of the Qx hexaphyrins in the lowest triplet state are calculated instead of
state was estimated via the subtraction of steady-state spectra, those in the lowest singlet excited state. For P26H, the averaged
including the ground-state absorption and fluorescence spectra NICS(0) value calculated at the peripheral positions is evaluated
from the DAS of the Qx state. Along the same lines, two DAS are as a negative value in the ground state, indicative of its aromatic
obtained, originating from the Qx and SD (optically dark state) character. On the other hand, a positive value is observed in the
states; S2 and S1 states with time constants of 700 fs and 16 ps, T1 state (Table 10). This result is in good agreement with the
respectively, are observed for P28H. Because of the absence of
fluorescence in P28H, the ESA spectra of both Qx and SD states Table 10. Calculated NICS Values at Various Positions of
are estimated by extraction from the ground-state absorption P26H and P28H in the Ground and Lowest Triplet States57
spectra and DAS.
P26H P28H
In porphyrin-related compounds, strong absorption peaks
observed in the ESA spectra can be assigned as transitions from position S0 T1 S0 T1
Qx to higher singlet excited states (Sn) with gerade symmetry NICS(0) a −15.15 10.77 14.81 −16.33
according to the selection rule for one-photon optical transition b −14.83 10.25 14.70 −16.31
because both B and Q states have Eu symmetry. In addition, average −14.99 10.51 14.76 −16.32
quantum chemical calculations of vertical excitation energies NICS(1) a −14.17 8.18 11.15 −15.15
afford numerous one-photon dark states with gerade symmetry b −13.17 7.42 11.06 −15.12
in the region above the B states in both P26H and P28H. These average −13.67 7.80 11.11 −15.14
energetic and symmetric relationships enable the assignment of NICS(2) a −12.12 4.26 6.53 −12.34
the observed ESA bands for P26H and P28H. For comparatively b −10.73 3.56 6.45 −12.22
analyzing the optical features between the ground and singlet average −11.43 3.91 6.49 −12.28
excited states, the extinction coefficients for the ground- and NICS(3) a −8.48 2.27 4.11 −8.55
singlet excited-state absorptions are calculated, as shown in b −7.80 1.95 4.06 −8.47
Figure 71.72 Notably, ESA from the Qx states of two hexaphyrins average −8.14 2.11 4.09 −8.51
is considered herein because the S1 state originating from the
direct HOMO−LUMO transition exhibits characteristics phenomena of aromaticity reversal changing from the ground
significantly different from the representative B and Q states in state to the T1 state. On the other hand, the opposite change of
porphyrinoids. By comparison with the ground-state absorption aromaticity is also observed in P28H. The NICS(0) value of
spectrum, the ESA spectrum of the Qx state in P26H is relatively P28H in the T1 state is negative, while that in the ground state is
broad and four times weaker in intensity, quite similar to the positive. As shown in Table 10, averaged NICS(1)−NICS(3)
typical absorption features observed for antiaromatic porphyr- values also demonstrate the (anti)aromaticity reversal of P26H
inoids. Hence, the lowest singlet excited state (Qx state) of P26H and P28H in the lowest triplet states as compared with that in
presumably becomes antiaromatic. On the other hand, the their ground states. Because the identical and similar magnitude
spectral shape and intensity of the ESA of the Qx state in P28H of the NICS values is computed at different sites, the NICS values
are sharpened and intensified, which could be assumed to be of P26H and P28H are considered to reflect the global
aromatic. These aspects of spectral changes are nearly similar to aromaticity, induced by the main π-conjugation pathway rather
those observed in previous bisrhodium hexaphyrins between the than by the local aromaticity triggered by local aromatic rings,
ground and lowest triplet states.52 Thus, the (anti)aromaticity of such as pyrrole or benzene.
phenylene-bridged hexaphyrins in the Qx state is clearly reversed The ACID plots are also consistent with the above-described
as compared with that in the ground state. results. The general directions of ring-current flow in the ACID
For addressing the issue on the reversed aromaticity of plots are inverted from clockwise (aromatic) to counterclockwise
hexaphyrins in the Qx state as compared with that in the ground (antiaromatic) in the S0 and T1 states of P26H, respectively.
state, the calculation of NICS52 values and ACID137 maps of Likewise, for P28H, the direction of the induced ring current in
(anti)aromatic hexaphyrins were attempted. In this study, the the ACID plots is reversed. From the calculated magnetic
aromatic indices in the lowest triplet state were analyzed instead aromaticity indices of hexaphyrins in the ground and triplet
of the Qx state. Previous results for annulenes, such as states, the aromaticity of hexaphyrins in the singlet excited state is
cyclooctatetraene and benzene, revealed that the reversal of believed to become reversed as compared with that of the closed-
aromaticity in the lowest triplet state also could be observed in shell singlet state.
the singlet excited state. As compared to those in the ground Furthermore, for supporting the reversed aromaticity of P26H
state, aromaticity indices, such as NICS values, magnetic and P28H in the singlet excited state (Qx state) as compared with
susceptibilities, and energy levels in the singlet excited state, that in the closed-shell singlet ground state, the HOMA values
are reversed, which is similar to those in the lowest triplet based on the ground- and excited-state optimized structures were
state.47,48,165−167 Furthermore, the experimental results obtained estimated (Table 11). For analyzing aromatic features according
for P26H and P28H clearly indicate reversed features in their Qx
states as compared with those in the ground state, which is similar Table 11. Evaluated HOMA Values of P26H and P28H in the
to the results obtained from the reversal of aromaticity in S0, T1, and Qx States for Main π-Conjugation Pathway57
rhodium hexaphyrins in the lowest triplet state.71 Accordingly,
the aromaticity of hexaphyrins in the singlet excited state is P26H P28H
assumed to be relevant to that in the lowest triplet state.
S0 T1 Qx S0 T1 Qx
Furthermore, it is difficult to calculate the singlet excited-state
properties of macrocyclic compounds, such as hexaphyrins, HOMA 0.80 0.71 0.78 0.60 0.79 0.76

2302 DOI: 10.1021/acs.chemrev.6b00313


Chem. Rev. 2017, 117, 2257−2312
Chemical Reviews Review

to the topological changes among S0, T1, and Qx states in more spectra provided strong evidence on the reversal of aromaticity in
detail, the HOMA values for the main [26] and [28]π- their singlet excited states. These far-reaching results provide us
conjugation pathways of P26H and P28H, respectively, were with useful information on the relationship between aromaticity
computed, as illustrated in Chart 9.72 As compared with the and electronic structures and, hence, the chemical reactivity and
HOMA values of the ground state, those of the T1 and Qx states stability of porphyrinoids in their excited states.
in P26H decrease, while those of the T1 and Qx states in P28H 5.3. Reversal of Aromaticity for Möbius Compounds in the
increase. Taking the similar tendency between T1 and Qx states Triplet Excited State
and the instability of antiaromatic molecules, which induces a Finally, in the last section, the T1-state aromaticity of Möbius
change to nonaromatic ones via structural distortions, into expanded porphyrins is illustrated for examining whether Baird’s
consideration, it could be suggested that the (anti)aromaticity of rule is applicable to Möbius annulenes.73,148 Certain expanded
P26H and P28H is reversed in their singlet excited states. In porphyrins are conformationally pliable to adapt half-twisted
addition, for clearly observing the reversal of aromaticity in the Möbius structures with the preservation of macrocyclic π-
singlet excited states of P26H and P28H as compared to their conjugation, enabling them to be suitable for the investigation of
ground states, MPD was evaluated (Figure 73).70,127 The MPD aromaticity; in addition, these expanded porphyrins can often
assume [4n]/[4n+2]π electronic states, which may enable a
comparative study of the excited states of Möbius aromatic and
antiaromatic congeners that have analogous molecular frame-
work. In addition, detailed analyses reveal that the spectroscopic
features of expanded porphyrins, i.e., absorption, fluorescence,
S1-state lifetime, and NLO properties, are attributed to their
[4n]/[4n+2]π-electronic states, providing obvious experimental
indices for determining excited-state aromaticity.38,39,53,168,169
In this regard, [28]hexaphyrin PdII complex Pd28H, N-fused
[22]pentaphyrin RhI complex Rh22P, and N-fused [24]-
pentaphyrin RhI complex Rh24P (Chart 10) were chosen,

Chart 10. Comparable Set of Metalated Expanded Porphyrins,


Pd28H, Rh22P, and Rh24P (Reprinted with Permission from
Ref 73; Copyright 2016 Wiley-VCH Verlag GmbH & Co.
KGaA)

Figure 73. MPD diagram of (a) P26H and (b) P28H in the ground and
singlet excited state on the basis of their optimized structures (B3LYP/
6-31G(d,p) level). Reprinted with permission from ref 57. Copyright
2015 American Chemical Society.

of P26H in the ground state exhibits a reduced value as compared


with that in the Qx state, supporting the reversal of aromaticity in
P26H in the Qx state due to structural distortion by
antiaromaticity. For P28H, on the other hand, MPD exhibits
an increased value in the ground state as compared with that
observed in the Qx state, which is also in good agreement with the
above results. That is, these results also indicate that the
aromaticity and antiaromaticity of hexaphyrins in the S0 state are
switched in their Qx state. With these MPD values, the dihedral
angles of bonds in the main π-conjugation pathways were
investigated. The mean values and standard deviations of
dihedral angles in the Qx state of P26H are increased in
comparison with those in the ground state. In contrast to P26H, where coordinated metals (palladium and rhodium) are
P28H exhibits decreased mean values and standard deviations supposed to accelerate intersystem crossing in their S1 states,
for the dihedral angles in the Qx state as compared to those in the affording corresponding T1 states. The previous research studies
ground state.57 In this regard, these aspects of the dihedral angle have studied that Rh22P exhibits Hückel aromatic characters,
changes in the Qx state were also observed in their T1 state. These and Pd28H and Rh24P show Möbius aromatic features based on
results also imply that hexaphyrins exhibit reversed aromatic experimental and theoretical results.170−172 Strong B-like bands
features in the Qx state as compared to the S0 state, representing and weak Q-like bands, features characteristic of aromatic
behavior similar to that observed in the T1 state. porphyrinoids, are observed in their ground-state absorption
In summary, the singlet excited-state behavior of Hückel spectra (Figure 74), also indicating that both Hückel and Möbius
aromatic and antiaromatic hexaphyrins observed from their TA aromatic expanded porphyrins exhibit similar π-electronic
2303 DOI: 10.1021/acs.chemrev.6b00313
Chem. Rev. 2017, 117, 2257−2312
Chemical Reviews Review

structures, in spite of their different molecular struc-


tures.37,62,64,110,173−175

Figure 74. S0- and T1-state absorption spectra of (a) Pd28H, (b) Rh22P
and (c) Rh24P. The asterisks indicate an experimental error due to the
different spectra resolutions between the two spectrometers used in
ground-state absorption and TA measurements. Reprinted with
permission from ref 73. Copyright 2016 Wiley-VCH Verlag GmbH &
Co. KGaA.
Figure 75. TA contour map (top) and decay-associated spectra
The aromatic character of Pd28H, Rh22P, and Rh24P has (bottom) of (a) Pd28H, (b) Rh22P, and (c) Rh24P in toluene.
been investigated by femtosecond TA spectra, which all exhibit Reprinted with permission from ref 73. Copyright 2016 Wiley-VCH
strong GSB and relatively weak ESA spectra (Figure 75) as Verlag GmbH & Co. KGaA.
aromatic characteristic features of aromatic porphyri-
noids.38,39,54,168,169 These TA spectra of Pd28H, Rh22P, and
Rh24P decay with two lifetime components, one in tens of decreased as compared with those of the corresponding S0
picosecond and the other longer than tens of nanoseconds, absorption spectra. Overall, these spectral characteristics of the
corresponding to the S1- and T1-state lifetimes, respectively. T1 states of Pd28H, Rh22P, and Rh24P indicate their
These TA spectra indicate that the intersystem crossing antiaromatic character.
processes to the T1 state are promoted because of the heavy These T1 absorption spectra allow the exploration of π-
atom effect of metal ions and become competitive with the rapid electronic structures in the T1 state, revealing their (anti)-
decay of the S1 state. aromaticity. In molecular systems, the absorption spectral
The global analysis of TA spectra by SVD can help in obtaining features have been well-established to reflect their characteristic
the DAS of the S1- and T1-state species. As the DAS of the T1 MO diagrams, π-electronic states, and MO struc-
state should consist of GSB and ESA spectra without SE, the T1 tures.38,39,54,168,169 The strong B- and weak Q-like bands of
absorption spectra of Pd28H, Rh22P, and Rh24P can be aromatic porphyrinoids originate from the configuration
evaluated by extracting the S0 absorption spectra and DAS interaction among four FMOs, while the broad, weak bands
(Figure 74).71,72 The estimated absorption spectrum of Pd28H with characteristic tailing in the NIR region of antiaromatic
in the T1 state is significantly different from the absorption porphyrinoids are described by configuration interaction of six
spectrum of Pd28H in the S0 state, showing broad and weak FMOs and one-photon forbidden transition, representing an
bands in the visible region with a tail in the NIR region, and its optically dark state. In this regard, the strong B- and weak Q-like
absorbance is significantly decreased as compared with that of the bands in the S0 absorption spectra of Pd28H, Rh22P, and Rh24P
S0 state.73 The spectral changes from the S0 state to the T1 state of are described by the vertical excitation energies between the four
Pd28H are comparable to those of TFM28H, which show FMOs (Figures 76), describing their aromatic character.55 The
switchable features between Möbius aromaticity and Hückel weak, broad, and tailing spectral features of Pd28H, Rh22P, and
antiaromaticity according to the ambient temperature.64 With Rh24P in the T1 states correspond well to simulated vertical
regard to the fact that Pd28H exhibits the same framework as excitation energies, where the lowest one-photon forbidden
that of bisrhodium hexaphyrins, except for central metal, its S0 transition is in good agreement with the spectroscopic features of
and T1 absorption spectra resemble those of R26H and R28H, typical antiaromatic porphyrinoids.39 Thus, the obtained
respectively.71 The absorption spectra of Rh22P and Rh24P in absorption spectra in the T1 state likely provide evidence for
the T1 state exhibit similar weak, broad, and tailing spectral defining the characteristic antiaromatic feature of Pd28H,
features, and their molar absorptivities are considerably Rh22P, and Rh24P.
2304 DOI: 10.1021/acs.chemrev.6b00313
Chem. Rev. 2017, 117, 2257−2312
Chemical Reviews Review

Figure 76. Vertical excitation energy plots calculated with TD-DFT


(column) compared with absorption spectra of (a) Pd28H, (b) Rh22P, Figure 77. Comparison of optimized structures of (a) Pd28H, (b)
and (c) Rh24P between the ground (black) and triplet (red) states Rh24P, and (c) Rh22P in the S0 (blue) and T1 (red) states. For clarity,
(B3LYP/LANL2DZP level). The magnified vertical excitation energy central metals, meso-pentafluorophenyl groups, and hydrogen atoms are
plots in the region of 900−2000 nm (inset). Reprinted with permission omitted (B3LYP/LANL2DZP level). Reprinted with permission from
from ref 73. Copyright 2016 Wiley-VCH Verlag GmbH & Co. KGaA. ref 73. Copyright 2016 Wiley-VCH Verlag GmbH & Co. KGaA.

0.30 and Wr = 0.73/0.70 for Pd28H in S0/T1 and Tw = 0.62/0.65


For the detailed analysis of the reversal of aromaticity in the T1 and Wr = 0.38/0.35 for Rh24P in S0/T1. As compared with that
state, theoretical aromaticity indices were investigated for in the T1 state, the Tw value in the S0 state is lower. Given that low
Pd28H, Rh22P, and Rh24P in the S0 and T1 states. The NICS Tw with large Wr describes less local twist and reinforced
values clearly revealed the reversal of aromaticity in the T1 conjugation but loss of smoother bending, the change in Tw
state.52 In other words, the negative NICS values at the center of values in the S0 and T1 states is in agreement with aromaticity
macrocycles are evaluated for the S0 states of Pd28H, Rh22P, reversal along with changes in the dihedral angles. In addition,
and Rh24P, respectively, while positive NICS values are the Tw values for Rh24P are even greater than those for Pd28H,
evaluated for the corresponding T1 states. attributed to larger twisting strain caused by the small size of
For probing structural changes during photoexcitation, the annulene and fused tripentacyclic part of Rh24P.
optimized structures of Pd28H, Rh22P, and Rh24P in the S0 and The ACID137 plots also obviously provide the evidence of the
T1 states were compared; the overlay of S0 and T1 structures switched aromaticity of Pd28H, Rh22P, and Rh24P in the T1
specifies that their topologies are more distorted in the T1 state state (Figure 79).73Pd28H, Rh22P, and Rh24P exhibit reversed
(Figure 77).156,176−178 For a precise comparison, dihedral angle ring-current flows in the T1 states compared to their
distribution analyzed by standard deviation, and average values of corresponding S0 states; clockwise ring currents in the S0 state
the S0 and T1 states, are measured because they provide explicit indicate their aromaticity, while counterclockwise ring currents
information about conformation and cyclic π-conjugation (Table in the T1 state well indicate antiaromatic characters. Collectively,
12). As compared with the dihedral angles of Pd28H, Rh22P, these sets of calculated aromaticity indices are consistent with the
and Rh24P in the S0 state, those in the T1 state are spread more results obtained from spectroscopic properties, evidently
widely, and their average values and standard deviations increase supporting the reversal of aromaticity in the T1 state, or
in the T1 state, which is in line with the aromatic feature in the S0 compliance with Baird’s rule, in stable Möbius aromatic systems.
state and antiaromatic feature in the T1 state. These character- To sum up, the T1 absorption spectra of Hückel or Möbius
istics are also confirmed in the MPD of Rh22P.72Rh22P exhibits expanded porphyrins are measured by femtosecond TA
a slightly smaller MPD value in the S0 state than in the T1 state spectroscopy, which permits the comparative studies of photo-
(Figure 78), supporting the reversal of aromaticity in the T1 state. physical properties between the S0 and T1 states. By quantum
Nevertheless, contrary to Rh22P, the Möbius structures of calculations, the detailed study of absorption spectra between the
Pd28H and Rh24P do not allow for a topological study with S0 and T1 states reveals that the aromaticity in the S0 state is
MPD. Hence, the linking numbers (Lk) of Pd28H and Rh24P switched to the antiaromaticity in the T1 state in Möbius
are calculated, which implies the structural concept of p-orbital aromatic molecules in addition to Hückel aromatic systems.
overlap in the distorted topology.88,179−181 Lk, a positive integer
corresponding to the number of half-twists in the molecule, 6. SUMMARY AND OUTLOOK
consists of twist (Tw) and writhe (Wr), Lk = Tw + Wr. Here, Tw In this Review, the control of aromaticity in expanded porphyrins
indicates the local structural twist and Wr expresses the global has been described by experimental and theoretical approaches.
property for long-range bending. The Möbius aromatic Pd28H The molecular structures of expanded porphyrins could be
and Rh24P exhibit Lk = 1 in the S0 and T1 states, while the Tw and modulated by decreasing the temperatures, inserting and
Wr values are different between the S0 and T1 states; Tw = 0.27/ removing protons, and regulating the polarity of solvents, all of
2305 DOI: 10.1021/acs.chemrev.6b00313
Chem. Rev. 2017, 117, 2257−2312
Chemical Reviews Review

Table 12. Dihedral Angle Data for Pd28H, Rh22P, and Rh24P in the S0 and T1 States73
Pd28H Rh22P Rh24P
dihedral angles S0 T1 S0 T1 S0 T1
range (Δ) 4.1−46.4° (42.3°) 1.9−59.7° (57.8°) 2.3−32.4° (30.1°) 3.4−40.0° (36.4°) 2.1−50.6° (48.5°) 2.1−61.4° (59.3°)
average 18.0° 19.0° 15.9° 16.9° 21.6° 22.9°
standard deviation 12.9° 16.2° 12.7° 13.0° 15.0° 18.5°

porphyrins, the control of temperature, protons, and solvent


environment has been observed to effectively eliminate intra-
molecular and intermolecular hydrogen bonds and smoothly
release the conformational constraints of expanded porphyrins,
leading to a preferred geometry for realizing more stable
electronic conjugation. The results discussed in this Review
suggest that the strategy for controlling aromaticity is promising
for the efficient realization of novel switchable π-conjugation
pathways as well as for the enhancement of aromaticity. The
control of aromaticity and molecular structure will provide
valuable information on the potential for synthesizing supra-
molecular multiple π-conjugated molecules and developing
molecular electronic devices. Furthermore, the elaborate
modulation of aromaticity proposes that highly conjugated,
flexible molecular systems, such as expanded porphyrins, are an
ideal system for controlling aromaticity.182
In addition, throughout these studies on the control of
aromaticity in expanded porphyrins, we would like to emphasize
the extension of these studies to the development of functional
materials and/or devices. For instance, as two representative
examples, we expect that (1) aromaticity control will aid in the
assessment of the stability or reactivity of (anti)aromatic
Figure 78. MPD diagram of Rh22P in the ground and triplet state based
compounds as well as the efficiency of some photochemical
on the optimized structures (B3LYP/LANL2DZP level). Reprinted
with permission from ref 73. Copyright 2016 Wiley-VCH Verlag GmbH reactions involving (anti)aromatic systems and that (2)
& Co. KGaA. antiaromatic compounds can be utilized in the development of
rapid saturable absorbers operating in the NIR region.
The switching of aromaticity in the ground and excited states,
as well as the different chemical nature of (anti)aromatic systems
induced thereof, will permit the qualitative estimation of the
reactivity of certain chemical reactions of interest. For instance,
when an aromatic molecule undergoes aromaticity reversal by
photoexcitation to be antiaromatic, it is speculated that the
relative instability induced by its enhanced biradical character
would accelerate the rate of photoinduced addition reactions
with olefins, the spectral characteristics of which will be detected
in the TA spectra. Thus, the relative reactivity of excited-state
chemical reactions involving (anti)aromatic compounds, such as
olefin metathesis and bridge-forming reactions in porphyrin-
based architectures, can be visualized. This will, in turn, permit
the design and outline of chemically feasible photoinduced
reaction routes by referencing the results discussed herein on the
evaluation of such reactivity, which is expected to provide some
Figure 79. ACID plots of (a) Pd28H, (b) Rh22P and (c) Rh24P in the fundamental insights into the fabrication of (anti)aromatic
ground (top) and triplet (bottom) states (B3LYP/LANL2DZP level). molecular architectures.
Reprinted with permission from ref 73. Copyright 2016 Wiley-VCH Considering that the excited-state absorption features of
Verlag GmbH & Co. KGaA. aromatic (antiaromatic) molecules resemble those of antiar-
omatic (aromatic) counterparts and this whole idea of
which can induce changes in aromaticity. Furthermore, the aromaticity reversal, expanded porphyrin derivatives exhibiting
aromaticity of expanded porphyrins in the ground state is antiaromaticity in their ground states can be used as precursors
reversed in the lowest triplet and first singlet excited states. In for developing saturable absorbers operating in the NIR region.
other words, various experimental methods for controlling the When antiaromatic compounds are excited by light, these
aromaticity of expanded porphyrins can be employed by compounds exhibit strong excited-state absorption features, as
changing their topologies and electronic states. predicted by theory and other spectroscopic evidence for
During the course of the sustained studies on the relationship aromaticity reversal, which allows them to be potential
between the aromaticity and molecular topology of expanded candidates for efficient saturable absorbers. These excited-state
2306 DOI: 10.1021/acs.chemrev.6b00313
Chem. Rev. 2017, 117, 2257−2312
Chemical Reviews Review

absorption structures will typically appear in the NIR region and of expanded porphyrins and the exciton dynamics in various conjugated
can be tuned by using compounds with various peripheral oligo-/polymers and molecular assemblies.
functional groups attached. Simultaneously, sufficiently short Jong Min Lim, born in 1982 in Seoul, Korea, attended Yonsei University
intrinsic excited-state lifetimes of antiaromatic compounds will from 2007 to 2013, where he did doctoral research focusing on
contribute to the realization of rapid saturable absorbers, which photophysical properties in expanded porphyrins and molecular
primarily function on the picosecond timescale in the NIR aggregates under the direction of Prof. and Dr. Dongho Kim. After
region. receiving his Ph.D. degree, he joined the research group of Prof. Philipp
Finally, these results will provide further insight into the Kukura as a postdoctoral fellow at the University of Oxford, where he
comprehension of the relationship among molecular structures, studied time-resolved Raman spectroscopy and microscopy.
spectroscopic properties, and aromaticity. Moreover, these
experimental findings are consistent with theoretical analyses, Min-Chul Yoon, born in 1975 in South Korea, received his B.S. (1998)
which predict a change in the aromaticity of expanded and M.S. (2000) degrees from the Department of Chemistry at Inha
porphyrins. On the basis of present results discussed herein, it University. Then he joined the Korea Research Institute of Standards
is suggested that the ability to manipulate (anti)aromaticity by and Science as an assistant researcher (2000−2001). From 2006, he
the selection of preferred topologies (Hückel and Möbius studied as a Ph.D. student in the Department of Chemistry at Yonsei
topologies) and electronic states (ground state and singlet and University and received his doctoral degree in 2010 under the
triplet excited states) will provide a guideline for modulating the supervision of Prof. Dongho Kim. During 2010−2012 he continued
molecular properties and reactivities of multiconjugated systems. his research career in in the spectroscopy laboratory for Functional π-
Further studies of the modulation events are expected to provide Electronic Systems as a postdoctoral associate. Since 2012 he has been a
additional abundant insight into the relationship between principal engineer in manufacturing and engineering team of Samsung
electronic structures and (anti)aromaticity. Finally, this line of Electronics Co., where he has developed metrology and inspection
investigation could lead to the development of rapid saturable technology for semiconductors.
absorbers and novel molecular photoswitchable devices. Dongho Kim was born in 1957 in Seoul, Korea. He received his B.S.
Assuming that this inference is applicable, it could have a (1980) from Seoul National University and Ph.D. (1984) from
profound effect on practical applications, which complement the Washington University. After postdoctoral research at Princeton
advances in theoretical comprehension, possibly provided by our University, he joined the Korea Research Institute of Standards and
studies. Science (1986). In 2000, he moved to Yonsei University as a Professor of
Chemistry. He received the Scientist of the Month Award (1999), the
AUTHOR INFORMATION Sigma-Aldrich Award (2005), the Korea Science Award in Chemistry
(2006), the Star Faculty Award (2006), and the National Science
Corresponding Author Achieving Excellence Award (2010) and was selected as the Underwood
*E-mail: dongho@yonsei.ac.kr. Professor at Yonsei University three times (2007, 2010, and 2013). He
ORCID was a recipient of the National Research Initiatives Program (1997−
2006) named as “Center for Ultrafast Optical Characteristics Control”.
Dongho Kim: 0000-0001-8668-2644 He was also a leader in the Center for Smart Nano-Conjugates through
Notes the World Class University Program (2008−2013). Since 2013, he has
been in charge of the Global Research Laboratory Project named as
The authors declare no competing financial interest. “Characterization of Exciton Dynamics in Functional Nanostructures”.
Biographies His research activity has been focused on the spectroscopic
investigations of various π-conjugated molecular systems such as
Young Mo Sung, born in 1988 in Cheonan, Korea, received his B.S. porphyrin, pyrene, perylenebisimide, thiophene, and their assemblies
degree (2010) from the Department of Chemistry at Yonsei University, with a particular interest in excitation energy and electron-transfer
where he obtained his doctoral degree in 2015 under the supervision of dynamics both in ensemble and at single-molecule level. He has co-
Prof. Dongho Kim. Since 2015 he has worked as a postdoctoral associate authored more than 520 peer-reviewed articles and about 20 reviews in
in Spectroscopy Laboratory for Functional π-Electronic Systems journals and book publication. He has served as an editorial board
(FPIES), studying the excited-state aromaticity of expanded porphyrins. member for Bulletin of the Korean Chemical Society, Journal of Porphyrins
Juwon Oh, born in 1986 in Daejeon, Korea, received his B.S. degree and Phthalocyanines, and Journal of Physical Chemistry.
(2013) in Chemistry from Yonsei University. He is working in the
Spectroscopy Laboratory for Functional π-Electronic Systems with Prof. ACKNOWLEDGMENTS
Dongho Kim. His research is focused on the photophysical properties of
donor−acceptor molecular systems and expanded porphyrins depend- First, we express our sincere gratitude to the colleagues with
ing on their aromaticity. whom we have been collaborating over the years on this subject,
in particular Prof. Atsuhiro Osuka (Kyoto University) and his
Won-Young Cha was born in 1985 in Seoul, Korea. He received his B.S. skilled graduate students and postdoctoral fellows. We also thank
degree (2011) from the Department of Chemistry at Yonsei University our former and current laboratory members for their tremendous
and now is a Ph.D. student in Prof. Dongho Kim’s laboratory. His contributions to the research topics over the years dealt with in
current work is focused on the synthesis and characterization of new this Review. The research subjects described in sections 2−4
macrocyclic conjugated systems that possess multiple conjugation were financially supported by the Global Research Laboratory
pathways in an organic molecule. (GRL, 2013K1A1A2A0205183) Program funded by the
Woojae Kim was born in 1990 in Seoul, Korea. He received his B.S. Ministry of Education, Science and Technology (MEST) of
degree (2014) from the Department of Chemistry at Yonsei University Korea. The research topic in section 5 was supported by
and now is a Ph.D. student in Prof. Dongho Kim’s laboratory. His Samsung Science and Technology Foundation under Project no.
current research is focused on the aromaticity change in the excited state SSTF-BA1402-10. The quantum calculations were performed
2307 DOI: 10.1021/acs.chemrev.6b00313
Chem. Rev. 2017, 117, 2257−2312
Chemical Reviews Review

using the supercomputing resources of the Korea Institute of (23) Sessler, J. L.; Davis, J. M. Sapphyrins: Versatile Anion Binding
Science and Technology Information (KISTI). Agents. Acc. Chem. Res. 2001, 34, 989−997.
(24) Nishino, H.; Kosaka, A.; Hembury, G. A.; Aoki, F.; Miyauchi, K.;
REFERENCES Shitomi, H.; Onuki, H.; Inoue, Y. Absolute Asymmetric Photoreactions
of Aliphatic Amino Acids by Circularly Polarized Synchrotron
(1) Sessler, J. L.; Weghorn, S. J. Tetrahedron Organic Chemistry Series,
Radiation: Critically pH-Dependent Photobehavior. J. Am. Chem. Soc.
Vol. 15; Pergamon Press: Oxford, U.K., 1997.
(2) Jasat, A.; Dolphin, D. Expanded Porphyrins and Their Heterologs. 2002, 124, 11618−11627.
Chem. Rev. 1997, 97, 2267−2340. (25) Naumovski, L.; Ramos, J.; Sirisawad, M.; Chen, J.; Thiemann, P.;
(3) Lash, T. D. Giant Porphyrinoids: From Figure Eights to Lecane, P.; Magda, D.; Wang, Z.; Cortez, C.; Boswell, G.; et al.
Nanomolecular Cavities. Angew. Chem., Int. Ed. 2000, 39, 1763−1767. Sapphyrins Induce Apoptosis in Hematopoietic Tumor-Derived Cell
(4) Furuta, H.; Maeda, H.; Osuka, A. Confusion, Inversion, and Lines and Show in Vivo Antitumor Activity. Mol. Cancer Ther. 2005, 4,
CreationA New Spring from Porphyrin Chemistry. Chem. Commun. 968−976.
2002, 17, 1795−1804. (26) Wei, W.-H.; Wang, Z.; Mizuno, T.; Cortez, C.; Fu, L.; Sirisawad,
(5) Sessler, J. L.; Seidel, D. Synthetic Expanded Porphyrin Chemistry. M.; Naumovski, L.; Magda, D.; Sessler, J. L. New Polyethyleneglycol-
Angew. Chem., Int. Ed. 2003, 42, 5134−5175. Functionalized Texaphyrins: Synthesis and in Vitro Biological Studies.
(6) Sessler, J. L.; Tomat, E. Transition-Metal Complexes of Expanded Dalton Trans. 2006, 16, 1934−1942.
Porphyrins. Acc. Chem. Res. 2007, 40, 371−379. (27) Evans, J. P.; Xu, F.; Sirisawad, M.; Miller, R.; Naumovski, L.; Ortiz
(7) Inokuma, Y.; Osuka, A. Subporphyrins: Emerging Contracted de Montellano, P. R. Motexafin Gadolinium-Induced Cell Death
Porphyrins with Aromatic 14π-Electronic Systems and Bowl-Shaped Correlates with Heme Oxygenase-1 Expression and Inhibition of P450
Structures: Rational and Unexpected Synthetic Routes. Dalton Trans. Reductase-Dependent Activities. Mol. Pharmacol. 2007, 71, 193−200.
2008, 19, 2517−2526. (28) Misra, R.; Kumar, R.; Chandrashekar, T. K.; Suresh, C. H.; Nag,
(8) Tanaka, T.; Sugita, T.; Tokuji, S.; Saito, S.; Osuka, A. Metal A.; Goswami, D. 22π Smaragdyrin Molecular Conjugates with Aromatic
Complexes of Chiral Möbius Aromatic [28]Hexaphyrin(1.1.1.1.1.1): Phenylacetylenes and Ferrocenes: Syntheses, Electrochemical, and
Enantiomeric Separation, Absolute Stereochemistry, and Asymmetric Photonic Properties. J. Am. Chem. Soc. 2006, 128, 16083−16091.
Synthesis. Angew. Chem., Int. Ed. 2010, 49, 6619−6621. (29) Zahedi Avval, F.; Berndt, C.; Pramanik, A.; Holmgren, A.
(9) Bauer, V. J.; Clive, D. L. J.; Dolphin, D.; Paine, J. B.; Harris, F. L.; Mechanism of Inhibition of Ribonucleotide Reductase with Motexafin
King, M. M.; Loder, J.; Wang, S. W. C.; Woodward, R. B. Sapphyrins: Gadolinium (MGd). Biochem. Biophys. Res. Commun. 2009, 379, 775−
Novel Aromatic Pentapyrrolic Macrocycles. J. Am. Chem. Soc. 1983, 105, 779.
6429−6436. (30) Srinivasan, A.; Anand, V. G. G.; Pushpan, S. K.; Chandrashekar, T.
(10) Woodward, R. B. Presentation at the Aromaticity Conference, K.; Sugiura, K.-I; Sakata, Y. Core modified meso-aryl sapphyrins and
Sheffield, U.K., 1966. rubyrins: structural and anion receptor properties. J. Chem. Soc., Perkin
(11) Saito, S.; Osuka, A. Expanded Porphyrins: Intriguing Structures, Trans. 2000, 2, 1788−1793.
Electronic Properties, and Reactivities. Angew. Chem., Int. Ed. 2011, 50, (31) Kumar, R.; Misra, R.; Chandrashekar, T. K.; Nag, A.; Goswami,
4342−4373. D.; Suresh, E.; Suresh, C. H. One-Pot Synthesis of Core-Modified
(12) Day, V. W.; Marks, T. J.; Wachter, W. A. Large Metal Ion- Rubyrin, Octaphyrin, and Dodecaphyrin: Characterization and Non-
Centered Templated Reactions. A Uranyl Complex of Cyclopentakis(2- linear Optical Properties. Eur. J. Org. Chem. 2007, 2007, 4552−4562.
iminoisoindoline). J. Am. Chem. Soc. 1975, 97, 4519−4527. (32) Chandrashekar, T. K.; Prabhuraja, V.; Gokulnath, S.;
(13) Rexhausen, H.; Gossauer, A. The Synthesis of a New 22π- Sabarinathan, R.; Srinivasan, A. Fused core-modified meso-aryl expanded
Electron Macrocyclic: Pentaphyrin. J. Chem. Soc., Chem. Commun. 1983, porphyrins. Chem. Commun. 2010, 46, 5915−5917.
275−275. (33) Karthik, G.; Sneha, M.; Raja, V. P.; Lim, J. M.; Kim, D.; Srinivasan,
(14) Berger, R. A.; LeGoff, E. The Synthesis of a 22π-Electron A.; Chandrashekar, T. K. Core-Modified meso-Aryl Hexaphyrins with an
Tetrapyrrolic Macrocycle. [1,3,1,3]Platyrin. Tetrahedron Lett. 1978, 19, Internal Thiophene Bridge: Structure, Aromaticity, and Photodynamics.
4225−4228.
Chem. - Eur. J. 2013, 19, 1886−1890.
(15) Sessler, J. L.; Weghorn, S. J.; Hiseada, Y.; Lynch, V. Hexaalkyl
(34) Karthik, A.; Srinivasan, A.; Chandrashekar, T. K. meso-Aryl Core-
Terpyrrole: A New Building Block for the Preparation of Expanded
Modified Fused Sapphyrins: Syntheses and Structural Diversity. Org.
Porphyrins. Chem. - Eur. J. 1995, 1, 56−67.
Lett. 2014, 16, 3472−3475.
(16) Sessler, J. L.; Cyr, M. J.; Lynch, V.; McGhee, E.; Ibers, J. A.
(35) Stępień, M.; Latos-Grażyński, L.; Sprutta, N.; Chwalisz, P.;
Synthetic and Structural Studies of Sapphyrin, a 22-π-Electron
Pentapyrrolic “Expanded Porphyrin. J. Am. Chem. Soc. 1990, 112, Szterenberg, L. Expanded Porphyrin with a Split Personality: A Hückel-
2810−2813. Möbius Aromaticity Switch. Angew. Chem., Int. Ed. 2007, 46, 7869−
(17) Sessler, J. L.; Weghorn, S. J.; Lynch, V.; Johnson, M. R. Turcasarin, 7873.
the Largest Expanded Porphyrin to Date. Angew. Chem., Int. Ed. Engl. (36) Szyszko, B.; Sprutta, N.; Chwalisz, P.; Stępień, M.; Latos-
1994, 33, 1509−1512. Grażyński, L. Hückel and Möbius Expanded para-Benziporphyrins:
(18) Sessler, J. L.; Murai, T.; Lynch, V.; Cyr, M. An “Expanded Synthesis and Aromaticity Switching. Chem. - Eur. J. 2014, 20, 1985−
Porphyrin”: The Synthesis and Structure of a New Aromatic 1997.
Pentadentate Ligand. J. Am. Chem. Soc. 1988, 110, 5586−5588. (37) Yoon, Z. S.; Osuka, A.; Kim, D. Möbius Aromaticity and
(19) Sessler, J. L.; Furuta, H.; Král, V. Phosphate anion chelation and Antiaromaticity in Expanded Porphyrins. Nat. Chem. 2009, 1, 113−122.
base-pairing. Design of receptors and carriers for nucleotides and (38) Shin, J.-Y.; Kim, K. S.; Yoon, M.-C.; Lim, J. M.; Yoon, Z. S.; Osuka,
nucleotide analogues. Supramol. Chem. 1993, 1, 209−220. A.; Kim, D. Aromaticity and Photophysical Properties of Various
(20) Sessler, J. L.; Hemmi, G.; Mody, T. D.; Murai, T.; Burrell, A.; Topology-Controlled Expanded Porphyrins. Chem. Soc. Rev. 2010, 39,
Young, S. W. Texaphyrins: Synthesis and Applications. Acc. Chem. Res. 2751−2767.
1994, 27, 43−50. (39) Cho, S.; Yoon, Z. S.; Kim, K. S.; Yoon, M. C.; Cho, D. G.; Sessler,
(21) Young, S. W.; Qing, F.; Harriman, A.; Sessler, J. L.; Dow, W. C.; J. L.; Kim, D. Defining Spectroscopic Features of Heteroannulenic
Mody, T. D.; Hemmi, W.; Hao, Y.; Miller, R. a. Gadolinium (III) Antiaromatic Porphyrinoids. J. Phys. Chem. Lett. 2010, 1, 895−900.
Texaphyrin: A Tumor Selective Radiation Sensitizer That Is Detectable (40) Garratt, P. J. Aromaticity; Wiley: New York, 1986; pp 1−19.
by MRI. Proc. Natl. Acad. Sci. U. S. A. 1996, 93, 6610−6615. (41) Krygowski, T. M.; Cyrański, M. K. Structural Aspects of
(22) Sessler, J. L.; Miller, R. A. Texaphyrins: New Drugs with Diverse Aromaticity. Chem. Rev. 2001, 101, 1385−1419.
Clinical Applications in Radiation and Photodynamic Therapy. Biochem. (42) Rzepa, H. S. Möbius Aromaticity and Delocalization. Chem. Rev.
Pharmacol. 2000, 59, 733−739. 2005, 105, 3697−3715.

2308 DOI: 10.1021/acs.chemrev.6b00313


Chem. Rev. 2017, 117, 2257−2312
Chemical Reviews Review

(43) Cyrański, M. K. Energetic Aspects of Cyclic Pi-Electron (pentafluorophenyl) [28]Hexaphyrins(1.1.1.1.1.1) into Möbius Struc-
Delocalization: Evaluation of the Methods of Estimating Aromatic tures. J. Phys. Chem. A 2009, 113, 4498−4506.
Stabilization Energies. Chem. Rev. 2005, 105, 3773−3811. (64) Yoon, M.-C.; Kim, P.; Yoo, H.; Shimizu, S.; Koide, T.; Tokuji, S.;
(44) Hückel, E. Zur Quantentheorie der Doppelbindung. Eur. Phys. J. A Saito, S.; Osuka, A.; Kim, D. Solvent- and Temperature-Dependent
1931, 70, 204−286. Conformational Changes between Hückel Antiaromatic and Möbius
(45) Heilbronner, E. Hückel Molecular Orbitals of Möbius-Type Aromatic Species in meso-Trifluoromethyl Substituted [28]-
Conformations of Annulenes. Tetrahedron Lett. 1964, 5, 1923−1928. Hexaphyrins. J. Phys. Chem. B 2011, 115, 14928−14937.
(46) Baird, N. C. Quantum Organic Photochemistry. II. Resonance (65) Yoon, M.-C.; Shin, J.-Y.; Lim, J. M.; Saito, S.; Yoneda, T.; Osuka,
and Aromaticity in the Lowest 3ππ* State of Cyclic Hydrocarbons. J. Am. A.; Kim, D. Solvent-Dependent Aromatic versus Antiaromatic
Chem. Soc. 1972, 94, 4941−4948. Conformational Switching in meso-(Heptakis)pentafluorophenyl [32]-
(47) Karadakov, P. B. Ground- and Excited-State Aromaticity and Heptaphyrin. Chem. - Eur. J. 2011, 17, 6707−6715.
Antiaromaticity in Benzene and Cyclobutadiene. J. Phys. Chem. A 2008, (66) Saito, S.; Shin, J. Y.; Lim, J. M.; Kim, K. S.; Kim, D.; Osuka, A.
112, 7303−7309. Protonation-Triggered Conformational Changes to Möbius Aromatic
(48) Karadakov, P. B. Aromaticity and Antiaromaticity in the Low- [32]heptaphyrins(1.1.1.1.1.1.1). Angew. Chem., Int. Ed. 2008, 47, 9657−
Lying Electronic States of Cyclooctatetraene. J. Phys. Chem. A 2008, 112, 9660.
12707−12713. (67) Shin, J.-Y.; Lim, J. M.; Yoon, Z. S.; Kim, K. S.; Yoon, M.-C.;
(49) Schleyer, P. v. R.; Pühlhofer, F. Recommendations for the Hiroto, S.; Shinokubo, H.; Shimizu, S.; Osuka, A.; Kim, D. Conforma-
Evaluation of Aromatic Stabilization Energies. Org. Lett. 2002, 4, 2873− tional Changes of meso-Aryl Substituted Expanded Porphyrins upon
2876. Protonation: Effects on Photophysical Properties and Aromaticity. J.
(50) Cyrański, M. K.; Schleyer, P. v. R.; Krygowski, T. M.; Jiao, H.; Phys. Chem. B 2009, 113, 5794−5802.
Hohlneicher, G. Facts and Artifacts about Aromatic Stability Estimation. (68) Lim, J. M.; Shin, J.-Y.; Tanaka, Y.; Saito, S.; Osuka, A.; Kim, D.
Tetrahedron 2003, 59, 1657−1665. Protonated [4n]π and [4n+2]π Octaphyrins Choose Their Möbius/
(51) Krygowski, T. M.; Cyrański, M. K. Two Sources of the Decrease Hückel Aromatic Topology. J. Am. Chem. Soc. 2010, 132, 3105−3114.
of Aromaticity: Bond Length Alternation and Bond Elongation. Part II. (69) Cha, W.-Y.; Yoneda, T.; Lee, S.; Lim, J. M.; Osuka, A.; Kim, D.
An Analysis Based on Geometry of the Singlet and Triplet States of 4nπ Deprotonation Induced Formation of Mö bius Aromatic [32]-
Annulenes: C4H4, C5H5+, C6H62+, C7H7−, C8H8, C9H9+. Tetrahedron heptaphyrins. Chem. Commun. 2014, 50, 548−550.
1999, 55, 11143−11148. (70) Cha, W.; Lim, J. M.; Yoon, M.-C.; Sung, Y. M.; Lee, B. S.;
(52) von Rague Schleyer, P.; Maerker, C.; Dransfeld, A.; Jiao, H.; van Katsumata, S.; Suzuki, M.; Mori, H.; Ikawa, Y.; Furuta, H.; et al.
Eikema Hommes, N. J. R. Nucleus-Independent Chemical Shifts: A Deprotonation-Induced Aromaticity Enhancement and New Con-
Simple and Efficient Aromaticity Probe. J. Am. Chem. Soc. 1996, 118, jugated Networks in meso-Hexakis(pentafluorophenyl)[26]hexaphyrin.
6317−6318. Chem. - Eur. J. 2012, 18, 15838−15844.
(53) Stępień, M.; Sprutta, N.; Latos-Grażyński, L. Figure Eights, (71) Sung, Y. M.; Yoon, M.-C.; Lim, J. M.; Rath, H.; Naoda, K.; Osuka,
Mö bius Bands, and More: Conformation and Aromaticity of A.; Kim, D. Reversal of Hückel (anti)aromaticity in the Lowest Triplet
Porphyrinoids. Angew. Chem., Int. Ed. 2011, 50, 4288−4340. States of Hexaphyrins and Spectroscopic Evidence for Baird’s Rule. Nat.
(54) Yoon, M.-C.; Cho, S.; Suzuki, M.; Osuka, A.; Kim, D. Aromatic Chem. 2015, 7, 418−422.
versus antiaromatic effect on photophysical properties of conforma- (72) Sung, Y. M.; Oh, J.; Naoda, K.; Lee, T.; Kim, W.; Lim, M.; Osuka,
tionally Locked trans-vinylene-bridged hexaphyrins. J. Am. Chem. Soc. A.; Kim, D. A Description of Vibrational Modes in Hexaphyrins:
2009, 131, 7360−7367. Understanding the Aromaticity Reversal in the Lowest Triplet State.
(55) Gouterman, M. Spectra of Porphyrins. J. Mol. Spectrosc. 1961, 6, Angew. Chem., Int. Ed. 2016, 55, 11930−11934.
138−163. (73) Oh, J.; Sung, Y. M.; Kim, W.; Mori, S.; Osuka, A.; Kim, D.
(56) Berera, R.; van Grondelle, R.; Kennis, J. T. M. Ultrafast Transient Aromaticity Reversal in the Lowest Excited Triplet State of Archetypical
Absorption Spectroscopy: Principles and Application to Photosynthetic Möbius Heteroannulenic Systems. Angew. Chem., Int. Ed. 2016, 55,
Systems. Photosynth. Res. 2009, 101, 105−118. 6487−6491.
(57) Sung, Y. M.; Oh, J.; Kim, W.; Mori, H.; Osuka, A.; Kim, D. (74) Hajgato, B.; Huzak, M.; Deleuze, M. S. Focal Point Analysis of the
Switching between Aromatic and Antiaromatic 1,3-Phenylene-Strapped Singlet−Triplet Energy Gap of Octacene and Larger Acenes. J. Phys.
[26]- and [28]Hexaphyrins upon Passage to the Singlet Excited State. J. Chem. A 2011, 115, 9282−9293.
Am. Chem. Soc. 2015, 137, 11856−11859. (75) Jung, Y. M.; Noda, I. New Approaches to Generalized Two-
(58) Yoneda, T.; Kim, T.; Soya, T.; Neya, S.; Oh, J.; Kim, D.; Osuka, A. Dimensional Correlation Spectroscopy and Its Applications. Appl.
Conformational Fixation of a Rectangular Antiaromatic [28]- Spectrosc. Rev. 2006, 41, 515−547.
Hexaphyrin Using Rationally Installed PeriPheral Straps. Chem. - Eur. (76) Mori, S.; Osuka, A. Aromatic and Antiaromatic Gold(III)
J. 2016, 22, 4413−4417. Hexaphyrins with Multiple Gold−Carbon Bonds. J. Am. Chem. Soc.
(59) Lee, H.; An, S.-Y.; Cho, M. Nonlinear Optical (NLO) Properties 2005, 127, 8030−8031.
of the Octupolar Molecule: Structure−Function Relationships and (77) Mori, S.; Kim, K. S.; Yoon, Z. S.; Noh, S. B.; Kim, D.; Osuka, A.
Solvent Effects. J. Phys. Chem. B 1999, 103, 4992−4996. Peripheral Fabrications of a Bis-Gold(III) Complex of [26]-
(60) Lee, W.-H.; Lee, H.; Kim, J.-A.; Choi, J.-H.; Cho, M.; Jeon, S.-J.; Hexaphyrin(1.1.1.1.1.1) and Aromatic versus Antiaromatic Effect on
Cho, B. R. Two-Photon Absorption and Nonlinear Optical Properties of Two-Photon Absorption Cross Section. J. Am. Chem. Soc. 2007, 129,
Octupolar Molecules. J. Am. Chem. Soc. 2001, 123, 10658−10667. 11344−11345.
(61) Meyers, F.; Marder, S. R.; Pierce, B. M.; Brédas, J. L. Electric Field (78) Rybtchinski, B.; Sinks, L. E.; Wasielewski, M. R. Photoinduced
Modulated Nonlinear Optical Properties of Donor-Acceptor Polyenes: Electron Transfer in Self-Assembled Dimers of 3-Fold Symmetric
Sum-Over-States Investigation of the Relationship between Molecular Donor−Acceptor Molecules Based on Perylene-3,4:9,10-Bis-
Polarizabilities (α, β, and γ) and Bond Length Alternation. J. Am. Chem. (dicarboximide). J. Phys. Chem. A 2004, 108, 7497−7505.
Soc. 1994, 116, 10703−10714. (79) Yoon, Z. S.; Kwon, J. H.; Yoon, M.; Koh, M. K.; Noh, S. B.; Sessler,
(62) Sankar, J.; Mori, S.; Saito, S.; Rath, H.; Suzuki, M.; Inokuma, Y.; J. L.; Lee, J. T.; Seidel, D.; Aguilar, A.; Shimizu, S.; et al. Nonlinear
Shinokubo, H.; Suk Kim, K.; Yoon, Z. S.; Shin, J.-Y.; et al. Unambiguous Optical Properties and Excited-State Dynamics of Highly Symmetric
Identification of Möbius Aromaticity for meso-Aryl-Substituted [28]- Expanded Porphyrins. J. Am. Chem. Soc. 2006, 128, 14128−14134.
Hexaphyrins(1.1.1.1.1.1). J. Am. Chem. Soc. 2008, 130, 13568−13579. (80) Yoon, Z. S.; Cho, D.-G.; Kim, K. S.; Sessler, J. L.; Kim, D.
(63) Kim, K. S.; Yoon, Z. S.; Ricks, A. B.; Shin, J.; Mori, S.; Sankar, J.; Nonlinear Optical Properties as a Guide to Aromaticity in Congeneric
Saito, S.; Jung, Y. M.; Wasielewski, M. R.; Osuka, A.; et al. Temperature- Pentapyrrolic Expanded Porphyrins: Pentaphyrin, Sapphyrin, Isosmar-
Dependent Conformational Change of meso-Hexakis- agdyrin, and Orangarin. J. Am. Chem. Soc. 2008, 130, 6930−6931.

2309 DOI: 10.1021/acs.chemrev.6b00313


Chem. Rev. 2017, 117, 2257−2312
Chemical Reviews Review

(81) Lament, B.; Dobkowski, J.; Sessler, J. L.; Weghorn, S. J.; Waluk, J. (102) Boyd, R. W. Nonlinear Optics; Academic Press: Amsterdam, The
Spectroscopy and Photophysics of a Highly Nonplanar Expanded Netherlands, 2008; Chapter 12.
Porphyrin: 4,9,13,18,22,27-Hexaethyl-5,8,14,17,23,26-Hexamethyl- (103) Drobizhev, M.; Stepanenko, Y.; Dzenis, Y.; Karotki, A.; Rebane,
2,11,20-Triphenylrosarin. Chem. - Eur. J. 1999, 5, 3039−3045. A.; Taylor, P. N.; Anderson, H. L. Extremely Strong Near-IR Two-
(82) Song, N.-W.; Cho, H.-S.; Yoon, M.-C.; Aratani, N.; Osuka, A.; Photon Absorption in Conjugated Porphyrin Dimers: Quantitative
Kim, D.-H. Energy Relaxation Dynamics of Excited Triplet States of Description with Three-Essential-States Model. J. Phys. Chem. B 2005,
Directly Linked Zn (II) Porphyrin Arrays. Bull. Korean Chem. Soc. 2002, 109, 7223−7236.
23, 271−276. (104) Drobizhev, M.; Makarov, N. S.; Hughes, T.; Rebane, A.
(83) Gentemann, S.; Medforth, C. J.; Forsyth, T. P.; Nurco, D. J.; Resonance Enhancement of Two-Photon Absorption in Fluorescent
Smith, K. M.; Fajer, J.; Holten, D. Photophysical Properties of Proteins. J. Phys. Chem. B 2007, 111, 14051−14054.
Conformationally Distorted Metal-Free Porphyrins. Investigation into (105) Drobizhev, M.; Karotki, A.; Kruk, M.; Mamardashvili, N. Z.;
the Deactivation Mechanisms of the Lowest Excited Singlet State. J. Am. Rebane, A. Drastic Enhancement of Two-Photon Absorption in
Chem. Soc. 1994, 116, 7363−7368. Porphyrins Associated with Symmetrical Electron-Accepting Substitu-
(84) Gentemann, S.; Nelson, N. Y.; Jaquinod, L.; Nurco, D. J.; Leung, tion. Chem. Phys. Lett. 2002, 361, 504−512.
S. H.; Medforth, C. J.; Smith, K. M.; Fajer, J.; Holten, D. Variations and (106) Karotki, A.; Drobizhev, M.; Kruk, M.; Spangler, C.; Nickel, E.;
Temperature Dependence of the Excited State Properties of Conforma- Mamardashvili, N.; Rebane, A. Enhancement of Two-Photon
tionally and Electronically Perturbed Zinc and Free Base Porphyrins. J. Absorption in Tetrapyrrolic Compounds. J. Opt. Soc. Am. B 2003, 20,
Phys. Chem. B 1997, 101, 1247−1254. 321−332.
(85) Song, H.; Cissell, J. A.; Vaid, T. P.; Holten, D. Photophysics of (107) Yoon, M.-C.; Misra, R.; Yoon, Z. S.; Kim, K. S.; Lim, J. M.;
Reduced Silicon Tetraphenylporphyrin. J. Phys. Chem. B 2007, 111, Chandrashekar, T. K.; Kim, D. Photophysical Properties of Core-
2138−2142. Modified Expanded Porphyrins: Nature of Aromaticity and Enhance-
(86) Chirvony, V. S.; van Hoek, A.; Galievsky, V. A.; Sazanovich, I. V.; ment of Ring Planarity. J. Phys. Chem. B 2008, 112, 6900−6905.
Schaafsma, T. J.; Holten, D. Comparative Study of the Photophysical (108) Shin, J. Y.; Furuta, H.; Yoza, K.; Igarashi, S.; Osuka, A. meso-Aryl-
Properties of Nonplanar Tetraphenylporphyrin and Octaethylporphyr- Substituted Expanded Porphyrins. J. Am. Chem. Soc. 2001, 123, 7190−
in Diacids. J. Phys. Chem. B 2000, 104, 9909−9917. 7191.
(87) Sheik-Bahae, M.; Said, A. A.; Wei, T.-H.; Hagan, D. J.; Van (109) Tanaka, Y.; Shin, J.-Y.; Osuka, A. Facile Synthesis of Large meso-
Stryland, E. W. Sensitive Measurement of Optical Nonlinearities Using a Pentafluorophenyl-Substituted Expanded Porphyrins. Eur. J. Org. Chem.
Single Beam. IEEE J. Quantum Electron. 1990, 26, 760−769. 2008, 2008, 1341−1349.
(88) Rzepa, H. S. Lemniscular Hexaphyrins as Examples of Aromatic (110) Tanaka, Y.; Saito, S.; Mori, S.; Aratani, N.; Shinokubo, H.;
and Antiaromatic Double-Twist Möbius Molecules. Org. Lett. 2008, 10, Shibata, N.; Higuchi, Y.; Yoon, Z. S.; Kim, K. S.; Noh, S. B.; et al.
949−952. Metalation of Expanded Porphyrins: A Chemical Trigger Used To
(89) Ajami, D.; Oeckler, O.; Simon, A.; Herges, R. Möbius Aromatic
Produce Molecular Twisting and Möbius Aromaticity. Angew. Chem.,
Synthesis of a Hydrocarbon. Nature 2003, 426, 819−821.
Int. Ed. 2008, 47, 681−684.
(90) Shimizu, S.; Aratani, N.; Osuka, A. meso-Trifluoromethyl-
(111) Aihara, J.; Horibe, H. Macrocyclic Aromaticity in Hückel and
Substituted Expanded Porphyrins. Chem. - Eur. J. 2006, 12, 4909−4918.
Möbius Conformers of Porphyrinoids. Org. Biomol. Chem. 2009, 7,
(91) Allan, C. S. M.; Rzepa, H. S. Chiral Aromaticities. AIM and ELF
1939−1943.
Critical Point and NICS Magnetic Analyses of Mö bius-Type
(112) Kang, S.; Hayashi, H.; Umeyama, T.; Matano, Y.; Tkachenko, N.
Aromaticity and Homoaromaticity in Lemniscular Annulenes and
V.; Lemmetyinen, H.; Imahori, H. meso-3,5-Bis(trifluoromethyl)phenyl-
Hexaphyrins. J. Org. Chem. 2008, 73, 6615−6622.
(92) Fliegl, H.; Sundholm, D.; Taubert, S.; Pichierri, F. Aromatic Substituted Expanded Porphyrins: Synthesis, Characterization, and
Pathways in Twisted Hexaphyrins. J. Phys. Chem. A 2010, 114, 7153− Optical, Electrochemical, and Photophysical Properties. Chem. - Asian J.
7161. 2008, 3, 2065−2074.
(93) Ahn, T. K.; Kwon, J. H.; Kim, D. Y.; Cho, D. W.; Jeong, D. H.; (113) Kim, K. S.; Lim, J. M.; Osuka, A.; Kim, D. Various Strategies for
Kim, S. K.; Suzuki, M.; Shimizu, S.; Osuka, A.; Kim, D. Comparative Highly-Efficient Two-Photon Absorption in Porphyrin Arrays. J.
Photophysics of [26]- and [28]Hexaphyrins(1.1.1.1.1.1): Large Two- Photochem. Photobiol., C 2008, 9, 13−28.
P ho t on A b s or pt i o n Cr o ss S e c t i o n o f A r o m a t i c [ 2 6 ] - (114) Kwon, J. H.; Ahn, T. K.; Yoon, M.-C.; Kim, D. Y.; Koh, M. K.;
Hexaphyrins(1.1.1.1.1.1). J. Am. Chem. Soc. 2005, 127, 12856−12861. Kim, D.; Furuta, H.; Suzuki, M.; Osuka, A. Comparative Photophysical
(94) Lim, J. M.; Yoon, Z. S.; Shin, J.-Y.; Kim, K. S.; Yoon, M.-C.; Kim, Properties of Free-Base, Bis-Zn(II), Bis-Cu(II), and Bis-Co(II) Doubly
D. The Photophysical Properties of Expanded Porphyrins: Relation- N-Confused Hexaphyrins(1.1.1.1.1.1). J. Phys. Chem. B 2006, 110,
ships between Aromaticity, Molecular Geometry and Non-Linear 11683−11690.
Optical Properties. Chem. Commun. 2008, 24, 261−273. (115) Turro, N. J. Modern Molecular Photochemistry; The Benjamin/
(95) Englman, R.; Jortner, J. The Energy Gap Law for Radiationless Cumming Publishing Co.: Menlo Park, 1978.
Transitions in Large Molecules. Mol. Phys. 1970, 18, 145−164. (116) Werner, A.; Michels, M.; Zander, L.; Lex, J.; Vogel, E. Figure
(96) Mitchell, R. H. Measuring Aromaticity by NMR. Chem. Rev. 2001, Eight” Cyclooctapyrroles: Enantiomeric Separation and Determination
101, 1301−1315. of the Absolute Configuration of a Binuclear Metal Complex. Angew.
(97) Jacobsen, N. NMR Spectroscopy Explained: Simplified Theory, Chem., Int. Ed. 1999, 38, 3650−3653.
Applications and Examples for Organic Chemistry and Structural Biology; (117) Lintuluoto, J. M.; Nakayama, K.; Setsune, J. Direct
Wiley-Interscience: New York, 2007. Determination of Absolute Configuration of Carboxylic Acids by
(98) Bryant, R. G. The NMR Time Scale. J. Chem. Educ. 1983, 60, 933. Cyclooctapyrrole. Chem. Commun. 2006, 33, 3492−3494.
(99) Walsh, S. Non-Linear Curve Fitting Using Microsoft Excel Solver. (118) Setsune, J.; Tsukajima, A.; Okazaki, N.; Lintuluoto, J. M.;
Talanta 1995, 42, 561−572. Lintuluoto, M. Enantioselective Induction of Helical Chirality in
(100) Hiroto, S.; Shinokubo, H.; Osuka, A. Porphyrin Synthesis in Cyclooctapyrroles by Metal-Complex Formation. Angew. Chem., Int.
Water Provides New Expanded Porphyrins with Direct Bipyrrole Ed. 2009, 48, 771−775.
Linkages: Isolation and Characterization of Two Heptaphyrins. J. Am. (119) Mori, M.; Okawa, T.; Iizuna, N.; Nakayama, K.; Lintuluoto, J.
Chem. Soc. 2006, 128, 6568−6569. M.; Setsune, J. Dynamic Figure Eight Loop Structure of meso-
(101) Shimizu, S.; Taniguchi, R.; Osuka, A. meso-Aryl-Substituted Tetraaryl[32]octaphyrins(1.0.1.0.1.0.1.0). J. Org. Chem. 2009, 74,
[26]Hexaphyrin(1.1.0.1.1.0) and [38]Nonaphyrin(1.1.0.1.1.0.1.1.0) 3579−3582.
from Oxidative Coupling of a Tripyrrane. Angew. Chem., Int. Ed. 2005, (120) Yoon, Z. S.; Noh, S. B.; Cho, D.-G.; Sessler, J. L.; Kim, D.
44, 2225−2229. Evaluation of Planarity and Aromaticity in Sapphyrin and Inverted

2310 DOI: 10.1021/acs.chemrev.6b00313


Chem. Rev. 2017, 117, 2257−2312
Chemical Reviews Review

Sapphyrin Using a Bidirectional NICS (Nucleus-Independent Chemical (142) Cha, W.-Y.; Soya, T.; Tanaka, T.; Mori, H.; Hong, Y.; Lee, S.;
Shift) Scan Method. Chem. Commun. 2007, 2, 2378−2380. P ark , K . H.; Os uk a , A .; Kim , D. Mu ltifa ce te d [36]-
(121) Burgner, R. P.; Ponte Goncalves, A. M. Position-Dependent octaphyrin(1.1.1.1.1.1.1.1): Deprotonation-Induced Switching among
Deuterium Isotope Effects On The Radiationless Decay Of Triplet State Nonaromatic, Möbius Aromatic, and Hückel Antiaromatic Species.
Of Porphyrin Free Bases. Chem. Phys. Lett. 1977, 46, 275−278. Chem. Commun. 2016, 52, 6076−6078.
(122) Kajii, Y.; Obi, K.; Tanaka, I.; Tobita, S. Isotope Effects on (143) Reddy, J. S.; Mandal, S.; Anand, V. G. Cyclic Oligofurans: One-
Radiationless Transitions from the Lowest Excited Singlet State of Pot Synthesis of 30π and 40π Expanded Porphyrinoids. Org. Lett. 2006,
Tetraphenylporphin. Chem. Phys. Lett. 1984, 111, 347−349. 8, 5541−5543.
(123) Sobolewski, A. L.; Gil, M.; Dobkowski, J.; Waluk, J. On the (144) Rodriguez, J.; Kirmaier, C.; Holten, D. Time-resolved and static
Origin of Radiationless Transitions in Porphycenes. J. Phys. Chem. A optical properties of vibrationally excited porphyrins. J. Chem. Phys.
2009, 113, 7714−7716. 1991, 94, 6020−6029.
(124) Waluk, J. Ground- and Excited-State Tautomerism in (145) Rodriguez, J.; Holten, D. Ultrafast vibrational dynamics of a
Porphycenes. Acc. Chem. Res. 2006, 39, 945−952. photoexcited metalloporphyrin. J. Chem. Phys. 1989, 91, 3525−3531.
(125) Dorough, G. D.; Miller, J. R.; Huennekens, F. M. Spectra of the (146) Minkin, V. I.; Glukhovtsev, M. N.; Simkin, B. Y. In Aromaticity
Metallo-Derivatives of α,β,γ,δ-Tetraphenylporphine. J. Am. Chem. Soc. and Antiaromaticity: Electronic and Structural Aspects; Wiley: New York,
1951, 73, 4315−4320. 1994.
(126) Ojadi, E. C. A.; Linschitz, H.; Gouterman, M.; Walter, R. I.; (147) Ottosson, H. Organic Photochemistry: Exciting Excited-State
Lindsey, J. S.; Wagner, R. W.; Droupadi, P. R.; Wang, W. Sequential Aromaticity. Nat. Chem. 2012, 4, 969−971.
Protonation of meso-(p-Dimethylamino)phenyl)porphyrins: Charge- (148) Rosenberg, M.; Dahlstrand, C.; Kilså, K.; Ottosson, H. Excited
Transfer Excited States Producing Hyperporphyrin. J. Phys. Chem. 1993, State Aromaticity and Antiaromaticity: Opportunities for Photophysical
97, 13192−13197. and Photochemical Rationalizations. Chem. Rev. 2014, 114, 5379−5425.
(127) Suzuki, M.; Osuka, A. Reversible Caterpillar-Motion like (149) Aihara, J.-i. Aromaticity-based theory of pericyclic reactions. Bull.
Isomerization in a N,N′-Dimethyl hexaphyrin(1.1.1.1.1.1) Induced by Chem. Soc. Jpn. 1978, 51, 1788−1792.
Two-Electron Oxidation or Reduction. Chem. Commun. 2005, 29, (150) Ilić, P.; Sinković, B.; Trinajstić, N. Topological resonance
3685−3687. energies of conjugated structures. Isr. J. Chem. 1980, 20, 258−269.
(128) Weissleder, R. A Clearer Vision for in Vivo Imaging. Nat. (151) Fratev, F.; Monev, V.; Janoschek, R. Ab Initio Study of
Biotechnol. 2001, 19, 316−317. Cyclobutadiene in Excited States: Optimized Geometries, Electronic
(129) Weissleder, R.; Ntziachristos, V. Shedding Light onto Live Transitions and Aromaticities. Tetrahedron 1982, 38, 2929−2932.
Molecular Targets. Nat. Med. 2003, 9, 123−128. (152) Gogonea, V.; von Rague Schleyer, P.; Schreiner, P. R.
(130) Kertesz, M.; Choi, C. H.; Yang, S. Conjugated Polymers and Consequences of Triplet Aromaticity in 4nπ-Electron Annulenes:
Aromaticity. Chem. Rev. 2005, 105, 3448−3481. Calculation of Magnetic Shieldings for Open-Shell Species. Angew.
(131) Hagfeldt, A.; Grätzel, M. Molecular Photovoltaics. Acc. Chem. Chem., Int. Ed. 1998, 37, 1945−1948.
Res. 2000, 33, 269−277. (153) Zilberg, S.; Haas, Y. Two-State Model of Antiaromaticity: The
(132) Ikeda, S.; Toganoh, M.; Easwaramoorthi, S.; Lim, J. M.; Kim, D.; Low Lying Singlet States. J. Phys. Chem. A 1998, 102, 10843−10850.
Furuta, H. Synthesis and Photophysical Properties of N-Fused (154) Fowler, P. W.; Steiner, E.; Jenneskens, L. W. Ring-Current
Tetraphenylporphyrin Derivatives: Near-Infrared Organic Dye of Aromaticity in Triplet States of 4n π Electron Monocycles. Chem. Phys.
[18]Annulenic Compounds. J. Org. Chem. 2010, 75, 8637−8649. Lett. 2003, 371, 719−723.
(133) Harris, D. C. Quantitative Chemical Analysis; Freeman: New (155) Villaume, S.; Fogarty, H. A.; Ottosson, H. Triplet-State
York, 2007; pp 407−408. Aromaticity of 4nπ-Electron Monocycles: Analysis of Bifurcation in
(134) Woller, E. K.; DiMagno, S. G. 2,3,7,8,12,13,17,18-Octafluoro- the π Contribution to the Electron Localization Function. Chem-
5,10,15,20-Tetraarylporphyrins and Their Zinc Complexes: First PhysChem 2008, 9, 257−264.
Spectroscopic, Electrochemical, and Structural Characterization of a (156) Feixas, F.; Vandenbussche, J.; Bultinck, P.; Matito, E.; Solà, M.
Perfluorinated Tetraarylmetalloporphyrin. J. Org. Chem. 1997, 62, Electron Delocalization and Aromaticity in Low-Lying Excited States of
1588−1593. Archetypal Organic Compounds. Phys. Chem. Chem. Phys. 2011, 13,
(135) He, G. S.; Tan, L.-S.; Zheng, Q.; Prasad, P. N. Multiphoton 20690−20703.
Absorbing Materials: Molecular Designs, Characterizations, and (157) Zhu, J.; An, K.; von Rague Schleyer, P. Evaluation of Triplet
Applications. Chem. Rev. 2008, 108, 1245−1330. Aromaticity by the Isomerization Stabilization Energy. Org. Lett. 2013,
(136) Pawlicki, M.; Collins, H. A.; Denning, R. G.; Anderson, H. L. 15, 2442−2445.
Two-Photon Absorption and the Design of Two-Photon Dyes. Angew. (158) An, K.; Zhu, J. Evaluation of Triplet Aromaticity by the Indene-
Chem., Int. Ed. 2009, 48, 3244−3266. Isoindene Isomerization Stabilization Energy Method. Eur. J. Org. Chem.
(137) Geuenich, D.; Hess, K.; Köhler, F.; Herges, R. Anisotropy of the 2014, 2014, 2764−2769.
Induced Current Density (ACID), a General Method To Quantify and (159) Rath, H.; Aratani, N.; Lim, J. M.; Lee, J. S.; Kim, D.; Shinokubo,
Visualize Electronic Delocalization. Chem. Rev. 2005, 105, 3758−3772. H.; Osuka, A. Bis-Rhodium Hexaphyrins: Metalation of [28]hexaphyrin
(138) Saito, S.; Osuka, A. N-Fusion Reaction Sequence of and a Smooth Hückel Aromatic−antiaromatic Interconversion. Chem.
Heptaphyrin(1.1.1.1.1.1.1): Singly, Doubly, and Quadruply N-Fused Commun. 2009, 45, 3762−3764.
Heptaphyrins. Chem. - Eur. J. 2006, 12, 9095−9102. (160) Kadish, K. M.; Smith, K. M.; Guilard, R. Handbook of porphyrin
(139) Tokuji, S.; Shin, J.; Kim, K. S.; Lim, J. M.; Youfu, K.; Saito, S.; science, Vol. 1−20; World Scientific: 2010.
Kim, D.; Osuka, A. Facile Formation of a Benzopyrane-Fused (161) Buehl, M.; Thiel, W.; Jiao, H.; Schleyer, P. V. R.; Saunders, M.;
[28]Hexaphyrin That Exhibits Distinct Möbius Aromaticity. J. Am. Anet, F. a. L. Helium and Lithium NMR Chemical Shifts of Endohedral
Chem. Soc. 2009, 131, 7240−7241. Fullerene Compounds: An Ab Initio Study. J. Am. Chem. Soc. 1994, 116,
(140) Inoue, M.; Osuka, A. Redox-Induced Palladium Migrations That 6005−6006.
Allow Reversible Topological Changes between Palladium(II) Com- (162) Herges, R. Topology in Chemistry: Designing Mö bius
plexes of Möbius Aromatic [28]Hexaphyrin and Hückel Aromatic Molecules. Chem. Rev. 2006, 106, 4820−4842.
[26]Hexaphyrin. Angew. Chem., Int. Ed. 2010, 49, 9488−9491. (163) Kataoka, M. Magnetic Susceptibility and Aromaticity in the
(141) Koide, T.; Youfu, K.; Saito, S.; Osuka, A. Multiple Conforma- Excited States of Benzene. J. Chem. Res. 2004, 2004, 573−574.
tional Changes of β-Tetraphenyl meso-Hexakis(pentafluorophenyl) (164) Mori, H.; Lim, J. M.; Kim, D.; Osuka, A. Modulation of Dual
Substituted [26] and [28]hexaphyrins(1.1.1.1.1.1). Chem. Commun. Electronic Circuits of [26]Hexaphyrins Using Internal Aromatic Straps.
2009, 40, 6047−6049. Angew. Chem., Int. Ed. 2013, 52, 12997−13001.

2311 DOI: 10.1021/acs.chemrev.6b00313


Chem. Rev. 2017, 117, 2257−2312
Chemical Reviews Review

(165) Ottosson, H.; Kilså, K.; Chajara, K.; Piqueras, M. C.; Crespo, R.;
Kato, H.; Muthas, D. Scope and Limitations of Baird’s Theory on Triplet
State Aromaticity: Application to the Tuning of Singlet−Triplet Energy
Gaps in Fulvenes. Chem. - Eur. J. 2007, 13, 6998−7005.
(166) Rosenberg, M.; Ottosson, H.; Kilså, K. Influence of Excited State
Aromaticity in the Lowest Excited Singlet States of Fulvene Derivatives.
Phys. Chem. Chem. Phys. 2011, 13, 12912−12919.
(167) Jorner, K.; Emanuelsson, R.; Dahlstrand, C.; Tong, H.;
Denisova, A. V.; Ottosson, H. Impact of Ground- and Excited-State
Aromaticity on Cyclopentadiene and Silole Excitation Energies and
Excited-State Polarities. Chem. - Eur. J. 2014, 20, 9295−9303.
(168) Ottosson, H.; Borbas, K. E. Aromaticity: A Light-Switched Yin
and Yang Pair. Nat. Chem. 2015, 7, 373−375.
(169) Szyszko, B.; Białońska, A.; Szterenberg, L.; Latos-Grażyński, L.
Phenanthriporphyrin: An Antiaromatic Aceneporphyrinoid as a Ligand
for a Hypervalent Organophosphorus(V) Moiety. Angew. Chem., Int. Ed.
2015, 54, 4932−4936.
(170) Mori, S.; Shimizu, S.; Taniguchi, R.; Osuka, A. Group 10 Metal
Complexes of meso-Aryl-Substituted [26]Hexaphyrins with a Metal−
Carbon Bond. Inorg. Chem. 2005, 44, 4127−4129.
(171) Mori, S.; Shin, J.-Y.; Shimizu, S.; Ishikawa, F.; Furuta, H.; Osuka,
A. N-Fused Pentaphyrins and Their Rhodium Complexes: Oxidation-
Induced Rhodium Rearrangement. Chem. - Eur. J. 2005, 11, 2417−2425.
(172) Park, J. K.; Yoon, Z. S.; Yoon, M.; Kim, K. S.; Mori, S.; Shin, J.;
Osuka, A.; Kim, D. Möbius Aromaticity in N-Fused [24]Pentaphyrin
upon Rh(I) Metalation. J. Am. Chem. Soc. 2008, 130, 1824−1825.
(173) Pacholska-Dudziak, E.; Skonieczny, J.; Pawlicki, M.;
Szterenberg, L.; Ciunik, Z.; Latos-Grażyński, L. Palladium Vacatapor-
phyrin Reveals Conformational Rearrangements Involving Hückel and
Möbius Macrocyclic Topologies. J. Am. Chem. Soc. 2008, 130, 6182−
6195.
(174) Higashino, T.; Lim, J. M.; Miura, T.; Saito, S.; Shin, J.-Y.; Kim,
D.; Osuka, A. Möbius Antiaromatic Bisphosphorus Complexes of
[30]Hexaphyrins. Angew. Chem., Int. Ed. 2010, 49, 4950−4954.
(175) Higashino, T.; Lee, B. S.; Lim, J. M.; Kim, D.; Osuka, A. A
Möbius Antiaromatic Complex as a Kinetically Controlled Product in
Phosphorus Insertion to a [32]Heptaphyrin. Angew. Chem., Int. Ed.
2012, 51, 13105−13108.
(176) Zhu, J.; Fogarty, H. A.; Möllerstedt, H.; Brink, M.; Ottosson, H.
Aromaticity Effects on the Profiles of the Lowest Triplet-State Potential-
Energy Surfaces for Rotation about the CC Bonds of Olefins with
Five-Membered Ring Substituents: An Example of the Impact of Baird’s
Rule. Chem. - Eur. J. 2013, 19, 10698−10707.
(177) Streifel, B. C.; Zafra, J. L.; Espejo, G. L.; Gómez-García, C. J.;
Casado, J.; Tovar, J. D. An Unusually Small Singlet-Triplet Gap in a
Quinoidal 1,6-Methano[10]annulene Resulting from Baird’s 4 N π-
Electron Triplet Stabilization. Angew. Chem., Int. Ed. 2015, 54, 5888−
5893.
(178) Papadakis, R.; Ottosson, H. The Excited State Antiaromatic
Benzene Ring: A Molecular Mr Hyde? Chem. Soc. Rev. 2015, 44, 6472−
6493.
(179) Rappaport, S. M.; Rzepa, H. S. Intrinsically Chiral Aromaticity.
Rules Incorporating Linking Number, Twist, and Writhe for Higher-
Twist Möbius Annulenes. J. Am. Chem. Soc. 2008, 130, 7613−7619.
(180) Wannere, C. S.; Rzepa, H. S.; Rinderspacher, B. C.; Paul, A.;
Allan, C. S. M.; Schaefer, H. F.; Schleyer, P. V. R. The Geometry and
Electronic Topology of Higher-Order Charged Möbius Annulenes. J.
Phys. Chem. A 2009, 113, 11619−11629.
(181) Fliegl, H.; Sundholm, D.; Pichierri, F. Aromatic Pathways in
Mono- and Bisphosphorous Singly Mö bius Twisted [28] and
[30]hexaphyrins. Phys. Chem. Chem. Phys. 2011, 13, 20659−20665.
(182) Oh, J.; Mori, H.; Sung, Y. M.; Kim, W.; Osuka, A.; Kim, D.
Switchable π-Electronic Network of Bis(α-Oligothienyl)-Substituted
Hexaphyrins between Helical versus Rectangular Circuit. Chem. Sci.
2016, 7, 2239−2245.

2312 DOI: 10.1021/acs.chemrev.6b00313


Chem. Rev. 2017, 117, 2257−2312

You might also like