Dynamic functions of bis- and tris(saloph) cobalt(III) structures based on axial coordination

Shigehisa Akine ab
aNano Life Science Institute (WPI-NanoLSI), Kanazawa University, Kakuma-machi, Kanazawa 920-1192, Japan. E-mail: akine@se.kanazawa-u.ac.jp
bGraduate School of Natural Science and Technology, Kanazawa University, Kakuma-machi, Kanazawa 920-1192, Japan

Received 25th June 2025 , Accepted 21st August 2025

First published on 30th August 2025


Abstract

This review article focuses on the functionalization and dynamic functional switching of low-spin d6 cobalt(III) complexes derived from various oligo(saloph) structures, such as bis(saloph) macrocycles and tris(saloph) cage complexes (H2saloph = N,N′-disalicylidene-o-phenylenediamine). The bis(saloph) dicobalt(III) complexes with methylene- or phenylene-bridged ligands exhibit reversible redox-driven structural switching, in which the axial functional ligands dissociate and reassociate in response to the CoIII/CoII interconversion. The ether-bridged macrocyclic bis(saloph) cobalt(III) complexes show excellent cation binding affinity at the central O6 binding site, which is significantly influenced by the nature of the axial ligands at the cobalt centers. In particular, an anion-capped structure leads to the formation of a unique metastable host–guest complex, enabling stimuli-responsive behavior upon external triggering. Post-metalation ligand exchange with anionic ligands and bridging diamine ligands provides a versatile strategy for structural and functional tuning of these macrocyclic hosts. In some complexes, the ligand exchange reactivity and the guest binding affinity enhance each other. A helical tris(saloph) cobalt(III) cryptand exhibits dynamic P/M chirality interconversion via the axial ligand exchange involving achiral or chiral amines, allowing precise control over the chirality inversion rates and enabling a unique transient chirality inversion during racemization. Furthermore, closed-cage metallocryptands bearing bridging diamine ligands effectively suppress the guest uptake/release kinetics. Thus, the introduction, removal, and exchange of axial ligands (X) in the [Co(saloph)X2]+-type units have been successfully employed for the functionalization and dynamic switching of metallohosts and metallo-supramolecular structures.


image file: d5dt01501d-p1.tif

Shigehisa Akine

Shigehisa Akine received his Ph.D. in 2000 from the University of Tokyo under the supervision of Prof. Takayuki Kawashima, then worked as a research associate with Prof. Tatsuya Nabeshima at the University of Tsukuba. He became assistant professor in 2004 and associate professor in 2008. In 2013, he joined Kanazawa University as a full professor and in 2017 became a principal investigator at the Nano Life Science Institute (WPI-NanoLSI). His research focuses on kinetic control of structural conversions in responsive molecules, especially in host–guest, supramolecular, coordination, and organic chemistry.


1. Introduction

The salen ligand1 and its phenylene analog, the saloph ligand,1 are versatile tetradentate chelate ligands that provide a planar N2O2 coordination environment (Fig. 1a). Some of the resulting transition metal complexes have been used as catalysts,2 liquid crystals,3 and sensors,4 and are known to exhibit unique magnetic5 and optical properties,6 as well as biological activities.7 In addition to simple monomeric structures, many oligomeric compounds containing multiple salen units have been developed,8,9 and their material applications10 and catalytic functions11 have been studied. These include macrocyclic oligomers,8,9 as well as acyclic oligomers12 and molecular cages.13 In particular, cyclic oligomers and cage structures are advantageous because they can be obtained in relatively high yields, thanks to the dynamic nature of the imine C[double bond, length as m-dash]N double bonds.14 Some of these compounds exhibit unique host–guest binding affinities toward molecular and ionic guest species,9 taking advantage of the rigid and shape-persistent nature of the salen/saloph complex structures, which arises from their well-defined chelate coordination motifs.
image file: d5dt01501d-f1.tif
Fig. 1 (a) Chemical structures of H2salen, H2saloph, and [Co(saloph)X2]+. Functional groups can be introduced as axial ligands X. (b) X/Y axial ligand exchange in [Co(saloph)X2]+ structures, allowing for tunable properties at the axial positions.

Among the metals that can be incorporated into these oligo(salen)/-(saloph) structures for metallohosts and metallosupramolecular systems, those with a d8 electron configuration, such as nickel(II)15 and palladium(II),16 are particularly advantageous. Their square-planar geometry aligns well with the planar N2O2 coordination environment provided by the salen/saloph ligands. Metalation of these ligands yields rigid square planar complexes through the simultaneous formation of four coordination bonds, which facilitates the predictable and selective formation of shape-persistent structures. In addition, the resulting d8 square-planar complexes are diamagnetic. The use of diamagnetic metals is, in fact, essential for the investigation of metal-containing supramolecular and host–guest structures, as it allows for NMR measurements for structural elucidation and detailed analysis of host–guest binding.

In contrast to the square-planar metal ions, pentacoordinate metal centers situated in the salen/saloph coordination pocket can accommodate an additional monodentate ligand. For example, parent salen/saloph ligands are known to form mononuclear pentacoordinate complexes formulated as [Zn(salen)X]/[Zn(saloph)X] (X = H2O, py, etc.)17 or dimeric complexes [Zn2(salen)2]/[Zn2(saloph)2] in which one of the phenoxo oxygen atoms occupies the apical position of the counterpart Zn2+.18 This fifth coordination has been utilized for the construction, structural conversion, and functionalization of various types of multi-metal self-assembled structures.19–21

Analogously, hexacoordinate complexes, [M(salen)X2]/[M(saloph)X2], can be obtained by introducing two additional monodentate ligands (X) to the metal centers in the salen/saloph coordination site. In fact, the salen/saloph ligands can accommodate various metal ions with an octahedral geometry, which is the most common and ubiquitous coordination structure adopted by a wide range of transition metal ions. In most cases, the N2O2 donor set of the salen/saloph ligands occupies four equatorial positions around the octahedral metal ion. Accordingly, the two X ligands in the [M(salen)X2]/[M(saloph)X2] complexes occupy trans positions to each other, located at the axial positions relative to the MN2O2 plane in the [M(salen)]/[M(saloph)] structures (Fig. 1a).22

In particular, among various salen/saloph complexes containing an octahedral metal ion, the cobalt(III) complexes, [Co(salen)X2]+/[Co(saloph)X2]+,23,24 offer significant advantages. Owing to the large ligand field splitting originating from the low-spin d6 electron configuration, diamagnetic complexes are usually obtained exclusively and predictably, which facilitates investigation based on NMR spectroscopy. In addition, these low-spin cobalt(III) complexes are generally inert and stable, which allows various types of site-selective functionalizations at the axial positions without loss of the central cobalt(III) ion from the salen/saloph structures. Ligand exchange occurs slowly and only at the two axial X positions in the [Co(salen)X2]+/[Co(saloph)X2]+ structures, on a timescale of minutes to hours,25 owing to their inert nature (Fig. 1b). This reactivity is useful for the slow, time-dependent control of functions in the multi-metal structures.26 Although kinetic inertness often hampers the integration of dynamic functions into metal complexes, [Co(salen)]/[Co(saloph)] structures offer a distinct advantage: their axial positions remain sufficiently reactive while the equatorial CoN2O2 core maintains structural integrity. This spatially controlled reactivity enables the rational design of dynamic, switchable systems based on inert cobalt(III) centers.

Cobalt(III) complexes are particularly suited for functionally relevant structural transformations via selective axial ligand exchange, even though other transition metal complexes such as Pd, Pt, and Fe also exhibit kinetic inertness. Furthermore, the redox activity of cobalt(III), especially the CoIII/CoII couple, offers additional opportunities for external control and switching of functions,27 which is less accessible in Pd, Pt, Zn, or Fe systems. Indeed, a wide range of cobalt(III)-based redox-responsive functional molecules and supramolecular architectures have been developed, demonstrating the broad versatility and growing importance of switchable metal-containing structures.28

This review article focuses on the functionalization and dynamic switching of the low-spin d6 cobalt(III) complexes derived from various kinds of oligo(saloph) structures, such as bis(saloph) ligands (H4L1, H4L2, H4L3, H4L4) and the tris(saloph) cage (H6L5) (Fig. 2). In particular, introduction, removal, and exchange of the axial ligands X in the [Co(salen)X2]+/[Co(saloph)X2]+ structures26 have been efficiently employed for functionalization and the dynamic functional switching of metallohosts and metallo-supramolecular structures incorporating cobalt(III) centers.


image file: d5dt01501d-f2.tif
Fig. 2 Chemical structures of bis(saloph) compounds, H4L1, H4L2, H4L3, H4L4, and tris(saloph) cage H6L5.

2. Functions of bis(saloph) cobalt(III) complexes based on axial coordination

2.1. Methylene-bridged macrocyclic bis(saloph) dicobalt(III) complexes for redox-driven structural switching

A bis(saloph) macrocyclic ligand, H4L1 (Fig. 2),29 in which two H2saloph motifs are connected by two methylene linkers to form a macrocyclic framework, was synthesized. A series of doubly bridged dicobalt(III) complexes, L1Co2(DAn)2, was prepared by reacting H4L1 with cobalt(II) acetate in the presence of various diamine ligands (DAn = DA1–DA6) under aerobic conditions (Fig. 3a).30 These complexes were characterized by spectroscopic methods and X-ray crystallography, revealing dinuclear structures featuring two diamagnetic cobalt(III) ions bridged by diamine ligands, such as alkanediamines DA1–DA4 with varying methylene chain lengths (Fig. 3a). The use of oligo(ether) diamines, DA5 and DA6, introduced potential cation binding sites into the dicobalt(III) macrocycle, analogous to crown ethers.
image file: d5dt01501d-f3.tif
Fig. 3 Redox switching of dinuclear complexes L1Co2(DAn)2. (a) The oxidized state with octahedral cobalt(III) ions. (b) The reduced state with tetracoordinate cobalt(II) ions.

While the cobalt(III) ions in the L1Co2(DAn)2 complex prefer a hexacoordinate octahedral geometry (Fig. 3a), reduction to cobalt(II) leads to a less coordinated geometry (Fig. 3b). This CoIII/CoII redox transformation can be exploited to switch between the bound and unbound states of the axial ligands at the cobalt centers. Electrochemical measurements exhibited redox waves with large peak separations, suggesting that the redox processes are accompanied by significant changes in coordination geometry.30 Specifically, the CoIII → CoII reduction occurs in the hexacoordinate state with bridging diamine ligands, whereas the CoII → CoIII oxidation occurs in a tetracoordinate state lacking the diamine ligands (Fig. 3b). The reversible structural transformation between these two states was further confirmed by electrolytic absorption spectroscopy as well as mass spectrometry.

Thus, the doubly bridged structure can be constructed and destructed by association/dissociation of the axially coordinating diamine ligands DAn in response to redox changes. This would significantly change the host–guest binding behavior in the polyether-based cavities of L1Co2(DA5)2 and L1Co2(DA6)2. However, these dinuclear cobalt(III) complexes did not exhibit any binding affinity for alkali metal ions such as Li+, Na+, or K+.29 This may be attributed to the positive charge of the dicationic L1Co2 core, which experiences strong electrostatic repulsion with cationic guests.

2.2. Phenylene-bridged bis(saloph) dicobalt(III) complexes for redox-driven structural switching

The acyclic bis(saloph) ligand, H4L2 (Fig. 2), in which two H2saloph units share a single phenylenediamine subunit, is known to form a series of dinuclear complexes where the two metal centers are electronically coupled.20,24,31 This H4L2 ligand was used to prepare the dinuclear cobalt(III) complex L2Co2(B1)4, which features four crown ether subunits at the axial positions of the cobalt(III) centers (Fig. 4).32 The complex was synthesized by the reaction of H4L2 with cobalt(II) acetate in the presence of 4-aminomethylbenzo-15-crown-5 (B1) under aerobic conditions.
image file: d5dt01501d-f4.tif
Fig. 4 Redox-driven structural conversion of crown-ether-functionalized complex L2Co2(B1)4. (a) Oxidized state and (b) reduced state. Only the oxidized form L2Co2(B1)4 can bind a K+ ion in a sandwich fashion, enabling redox-switchable guest binding.

Since each face of the L2Co2(B1)4 complex functions as a bis(15-crown-5) host capable of binding a cationic guest in a sandwich fashion, the complex was expected to bind two K+ ions on both faces to form a 1[thin space (1/6-em)]:[thin space (1/6-em)]2 (host/guest) complex. However, spectroscopic measurements revealed that this L2Co2(B1)4 complex exhibits K+ binding with 1[thin space (1/6-em)]:[thin space (1/6-em)]1 stoichiometry, as confirmed by Job plot analysis and mass spectrometry. This binding behavior can be rationalized by the molecular deformation caused by the first guest binding in a [(15-crown-5)2K]+ sandwich fashion, which may increase the distance between the two crown ether moieties on the opposite face, thereby suppressing the second K+ binding. This L2Co2(B1)4 complex also undergoes a reversible CoIII/CoII redox interconversion, accompanied by the association/dissociation of the axially coordinating crown ether subunits B1 (Fig. 4).32

The CoIII/CoII redox interconversion of the same dinuclear L2Co2 motif was also exploited for the reversible association/dissociation of dendrimer subunits G3. The dinuclear complex L2Co2(G3)4, which contains four dendrimer subunits G3 (Fig. 5),33 was expected to undergo CoIII/CoII redox switching accompanied by the reversible binding of the dendrimer subunits G3. However, under electrochemical conditions, L2Co2(G3)4 did not exhibit efficient redox switching, likely due to the steric bulkiness of the dendrimer subunits G3. Instead, photo-driven redox reactions proved effective: upon photoirradiation (λ ≥ 420 nm) in a degassed DMF solution containing triethanolamine as a sacrificial electron donor, the complex was reduced to a tetracoordinate dicobalt(II) species. This reduced form can be re-oxidized by air exposure to regenerate the initial hexacoordinate dicobalt(III) complex. Thus, despite the steric hindrance from the bulky dendrimer subunits, reversible association/dissociation of the dendrimer subunits in L2Co2(G3)4 was successfully achieved via photo-driven redox switching.


image file: d5dt01501d-f5.tif
Fig. 5 Chemical structure of the dendrimer-functionalized dinuclear complex L2Co2(G3)4.

2.3. Ether-bridged bis(saloph) dicobalt(III) macrocycles for controlled guest binding

An ether-bridged bis(saloph) macrocycle, H4L3 (Fig. 2), which is an oxygen analogue of the methylene-bridged bis(saloph) macrocycle H4L1, was synthesized.34 Owing to the ether linkages, the metalated form, L3M2, features an 18-crown-6-like central binding cavity surrounded by six oxygen donor atoms, which exhibit excellent binding affinity for cationic guest species. For example, the nickel(II) metallohost L3Ni2 (Fig. 6a) strongly binds to Na+ to form a 1[thin space (1/6-em)]:[thin space (1/6-em)]1 host–guest complex and interacts with larger alkali metal ions (K+, Rb+, Cs+) to afford unique stacked structures.34–38 This higher binding affinity arises from the combination of negatively polarized phenoxo oxygen atoms and the well pre-organized structure of the metallohosts. This section focuses on the guest binding behavior of the hexacoordinate cobalt(III) analogues L3Co2A4 (Fig. 6b), which have four primary amine ligands A at the axial positions of the cobalt(III) centers.
image file: d5dt01501d-f6.tif
Fig. 6 Chemical structures of (a) L3Ni2 and (b) L3Co2A4.

Since the dinuclear cobalt(III) metallohosts L3Co2A4 are dicationic, they were expected to show a poor cation binding affinity due to electrostatic repulsion, especially in comparison to the non-charged nickel(II) analogue L3Ni2. Contrary to expectations, however, the cobalt(III) metallohost L3Co2(MeNH2)4, bearing four methylamine ligands, showed excellent binding affinity toward various cationic guests, such as a monovalent cation, Na+ (Ka = 8.5 × 106 M−1), and even a multivalent cation, La3+ (Ka = 2.4 × 106 M−1).39 In the crystal structure of the Na+ inclusion complex, L3Co2(MeNH2)4Na, the Na+ ion is located precisely at the center of the O6 binding site (Fig. 7a). The triflate counter anions not only directly coordinate to the Na+ ion but also form hydrogen bonds with the methylamine NH2 groups on both faces of the macrocycle, resulting in a unique anion-capped structure (Fig. 7b).39 These noncovalent interactions appear to contribute not only to the strong cation binding, but also to blocking guest entry/exit, acting like a cap or lid on a container.


image file: d5dt01501d-f7.tif
Fig. 7 (a) X-ray crystal structure of L3Co2(MeNH2)4Na. The two TfO counter anions are also shown. (b) Schematic drawing of the anion-capped structure of the host–guest complex, L3Co2(MeNH2)4Na. (c) Plots of mole fractions of the La3+ complex versus time after the addition of K+ in CD3OD: (i) in the absence of AcO; (ii) in the presence of AcO; (iii) guest exchange was initiated when AcO was added as a trigger after 120 h. (d) Conversion of the metastable state [K+ complex + unbound La3+] into the thermodynamically most stable state [La3+ complex + unbound K+] triggered by AcO. Adapted in part with permission from ref. 39.

In fact, the guest uptake/release rates of the metallohost L3Co2(MeNH2)4 were found to be slow on the 1H NMR time scale for cationic guests such as Na+, K+, and Ca2+. Notably, the uptake of the La3+ ion was particularly slow (kin ≈ 10−2 M−1 s−1), requiring nearly 100 h to reach completion. This remarkably slow uptake was attributed to the anion-capped structure, as the uptake rate was found to depend significantly on the nature of counter anions.

This anion-capped structure also contributed to the deceleration of guest exchange in the metallohost L3Co2(MeNH2)4. A compelling demonstration is the formation of a metastable host–guest complex, that is, a kinetically trapped state in which a weaker guest is preferentially bound within the host cavity, leaving a stronger guest unbound. From a thermodynamic viewpoint, the metallohost showed a lower affinity for K+ (Ka = 1.1 × 106 M−1) than for La3+ (Ka = 2.4 × 106 M−1). However, when K+ and La3+ were simultaneously added, K+ was selectively taken up for kinetic reasons. Guest exchange to the thermodynamically favored La3+ complex was not observed even after 2 weeks (Fig. 7c(i)), indicating that the guest exchange was almost completely kinetically suppressed. Typically, guest binding in simple crown ethers is fast enough that the thermodynamically most stable complex is always selectively formed. In this context, the cobalt(III) metallohost L3Co2(MeNH2)4 gives the first metastable host–guest complex derived from a simple macrocyclic host (Fig. 7d).39,40

Such a metastable state retains the potential to transition to the thermodynamically most stable structure at any time when triggered. In the aforementioned case, the mixture of the K+ complex with unbound La3+ remained in a kinetically trapped metastable state, where the conversion to the thermodynamically most stable state was suppressed, but this conversion was accelerated by the addition of acetate ion as a trigger (Fig. 7c(ii), (iii) and d). This represents a new type of on-demand, stimuli-responsive function that exploits the long-lived yet transformable nature of a metastable host–guest complex.

Thus, the guest uptake/release rates of the metallohost L3Co2(MeNH2)4 were found to be significantly decelerated by its anion-capped structure. These rates were also expected to be influenced by the structure of the amine ligands A coordinating to the cobalt(III) ions. In order to clarify this effect, a series of dinuclear metallohosts L3Co2A4, each bearing four primary monoamine ligands (A = EtNH2, PhCH2NH2, PhC2H4NH2, PhC3H6NH2) with or without a phenyl group remote from the O6 binding cavity, were synthesized (Fig. 6b). All these metallohosts exhibited a consistent selectivity trend among alkali metal ions: Na+ > K+ > Rb+ > Cs+. This trend was primarily attributed to differences of up to ∼500[thin space (1/6-em)]000-fold in the release rate constants kout, following the order of Na+ < K+ < Rb+ < Cs+.41

The structural variation in the amine ligands A also affected the binding behavior in both thermodynamic and kinetic aspects (Table 1). For example, the binding constants of L3Co2A4 (A = EtNH2, PhCH2NH2, PhC2H4NH2, PhC3H6NH2) with Na+ differed by up to 200-fold. The introduction of a phenyl group generally weakened the binding, with the benzylamine analogue, L3Co2(PhCH2NH2)4, showing the lowest affinity. A detailed analysis of the uptake/release rates, kin and kout, revealed that the phenyl-containing derivatives generally exhibited a faster release rate kout and slower uptake rate kin. Since the Na+-bound species of these phenyl-containing complexes, L3Co2(PhCnH2nNH2)4Na (n = 1,2,3), have quite similar structures to each other, in which the Na+ guest is located at the center of the O6 cavity, the differences in the binding constants Ka are mainly ascribed to variations in the uptake rates kin. Crystallographic analysis of the guest-free metallohosts, L3Co2(PhCnH2nNH2)4 (n = 1,2,3), revealed that some of the phenyl and methylene C–H groups interact with the oxygen atoms of the O6 binding site via C–H⋯O interactions. This suggests that guest binding requires additional energy to overcome this extra stabilization (Fig. 8).41


image file: d5dt01501d-f8.tif
Fig. 8 Guest binding equilibrium of L3Co2(PhCnH2nNH2)4 (n = 1,2,3). The definitions of the uptake and release kinetic rate constants, kin and kout, for the host–guest binding equilibrium are shown. Extra stabilization via C–H⋯O interactions by the phenyl capping can explain the relatively weaker binding compared to L3Co2(EtNH2)4.
Table 1 Binding constants and uptake/release rate constants for the host–guest complexes of L3Co2A4 with Na+[thin space (1/6-em)]a,b
  Ka (M−1) kin (M−1s−1) kout (s−1)
a In CD3OD.b Determined by 1H NMR spectroscopy. Data taken from ref. 41.
A = EtNH2 9.6 × 106 1.1 × 106 0.12
A = PhCH2NH2 4.9 × 104 5.2 × 104 1.05
A = PhC2H4NH2 3.8 × 105 9.0 × 104 0.23
A = PhC3H6NH2 8.1 × 105 1.0 × 105 0.12


2.4. Ligand exchange of bis(saloph) cobalt(III) metallohosts for control of the guest binding affinity

Since cobalt(III) is inert due to its low-spin d6 electron configuration, dissociation of the cobalt(III) ion from the tetradentate saloph chelate ligand is essentially negligible. Nevertheless, the Co–N bonds to the monoamine ligands A are relatively dynamic, thus allowing the ligand exchange in a site-selective fashion (Fig. 1b). This reactivity is advantageous for functionalization after the L3Co2 dinuclear structures are constructed, i.e., the post-metalation modification enables alteration or fine-tuning of molecular functions, such as the host–guest binding affinity, as seen in the post-synthetic modification of various molecular cages and MOFs.42 Indeed, the guest binding affinity of the dinuclear metallohosts L3Co2A4 was significantly influenced by the structure of the primary amine ligands A introduced at the cobalt(III) centers, as described in the previous section.

In this context, the synthesis of analogous L3Co2A4 complexes bearing different types of amine ligands, such as a secondary amine (pip = piperidine) and a tertiary amine (quin = quinuclidine), was attempted using the same protocol as that for L3Co2(MeNH2)4.43 Whereas the piperidine-coordinating complex, L3Co2(pip)4, was successfully obtained, the quinuclidine-coordinating complex, L3Co2(quin)4, was not obtained.

Regarding the ligand exchange reactivity, the methylamine ligands in L3Co2(MeNH2)4 were not readily exchanged with other amines (pip, quin) (Fig. 9a). In contrast, the piperidine-coordinating complex, L3Co2(pip)4, was efficiently converted into the methylamine-coordinating complex, L3Co2(MeNH2)4 (Fig. 9b), but not into the quinuclidine complex, L3Co2(quin)4 (Fig. 9c). Thus, the affinity order of the amines for the cobalt(III) centers in L3Co2A4 followed the trend of primary amine > secondary amine > tertiary amine, which can be attributed to the steric bulk around the nitrogen donor atom. Among the isolable L3Co2A4 complexes, the piperidine-coordinating complex, L3Co2(pip)4, was found to be the best starting complex for ligand exchange, as it exhibited the highest reactivity.43


image file: d5dt01501d-f9.tif
Fig. 9 Structural conversion of L3Co2(pip)4 and related complexes via ligand exchange. Path (h) represents the acceleration of ligand exchange with AcO in the presence of Na+.

It is noteworthy that the ligand exchange of L3Co2(pip)4 with pyridine selectively afforded a di-exchanged product, L3Co2(pip)2(py)2, instead of the tetra-exchanged species, L3Co2(py)4 (Fig. 9d). X-ray crystallography and 1H NMR spectroscopy revealed the anti-diagonal stereoconfiguration of this product, in which each cobalt center bears one pip and one py ligand occupying opposite positions on the two faces, e.g., pip above and py below on one cobalt center, and reverse on the other. This stereoselectivity can be attributed to the doubly curved structure of the L3Co2 framework, in which the bulkier piperidine ligands preferentially occupy the two convex faces, while the pyridine ligands reside on the narrower concave faces. The same product, the anti-diagonal L3Co2(pip)2(py)2, was also selectively formed via the reverse ligand exchange (py → pip) starting from L3Co2(py)4 (Fig. 9e).

A similar anti-diagonal di-exchanged complex, L3Co2(pip)2(OAc)2, was obtained by the reaction of L3Co2(pip)4 with acetate ion (Fig. 9f). This complex was obtained as precipitates directly from the reaction mixture in 82% isolated yield.

These results demonstrate that a variety of post-metalation modifications of L3Co2A4 metallohosts can be achieved by exploiting the reactivity of the axial ligands coordinating to the cobalt(III) centers.43 Such modulations provide a means to tune the binding affinity for cationic guests such as Na+ within the central O6 cavity.

The di-exchanged complex, L3Co2(pip)2(OAc)2, showed high affinity for cationic guests, owing to charge compensation of the dicationic L3Co2 core by the newly introduced anionic acetato ligands. Indeed, Na+ ion was quantitatively taken up to form the inclusion complex L3Co2(pip)2(OAc)2Na (Fig. 9g), whose structure was unambiguously determined by X-ray crystallography.

In contrast, when NaOTf was added to a CD3CN solution of the starting piperidine-coordinating complex, L3Co2(pip)4, formation of the Na+ inclusion complex was not observed, indicating its lower guest binding affinity. However, upon treatment of this mixture with AcO, ligand exchange rapidly proceeded, concomitant with Na+ uptake, to afford L3Co2(pip)2(OAc)2Na (Fig. 9h). These observations clearly demonstrate that the guest binding affinity of the piperidine-coordinating complex L3Co2(pip)4 is enhanced by exchange of the neutral piperidine ligands with anionic acetato ligands.43 Moreover, the presence of Na+ ion significantly enhanced the ligand exchange reactivity of the L3Co2(pip)4 metallohost.

Thus, the ligand exchange reactivity and the guest binding affinity of L3Co2(pip)4 were mutually enhanced. This interplay became even more evident in the guest binding of L3Co2(pip)4 accompanied by exchange with methoxo ligands under solvolytic conditions.

The piperidine-coordinating complex L3Co2(pip)4 slowly underwent solvolysis in CD3OD, affording a new species rather than immediately forming a simple guest-bound complex, L3Co2(pip)4Na.44 Spectroscopic and crystallographic investigations revealed that the resulting product was an inclusion complex L3Co2(pip)2(OMe)2Na, in which two piperidine ligands were replaced by methoxo ligands. Detailed analysis of the reaction progress indicated that the reaction first produced a mono-exchanged guest-bound species, L3Co2(pip)3(OMe)Na, which was then converted into the di-exchanged species, L3Co2(pip)2(OMe)2Na. In the first step, ligand exchange and Na+ uptake appear to occur concurrently.

More precisely, this process can be interpreted in terms of two possible mechanisms. One is the reaction first mechanism (Fig. 10A), in which exchange with the methoxo ligand occurs prior to guest binding. In this mechanism, the coordination of the anionic methoxo ligand cancels the positive charge of the cobalt(III) center, thereby enhancing the binding affinity for cationic guests in the central O6 cavity. The other is the recognition first mechanism (Fig. 10B), in which guest binding occurs before ligand exchange. In this mechanism, the presence of the guest in the central O6 cavity facilitates ligand exchange reaction at the cobalt(III) centers.


image file: d5dt01501d-f10.tif
Fig. 10 Guest uptake behavior of L3Co2(pip)4 associated with the ligand exchange reaction under solvolytic conditions. The recognition first mechanism (B) is suggested when Na+ is bound.

These two mechanisms can be distinguished by analyzing the ligand exchange kinetics at varying guest concentrations.45 In practice, the ligand exchange rate of L3Co2(pip)4 increased proportionally with the concentration of Na+, supporting the recognition first mechanism (Fig. 10B). In contrast, the ligand exchange rates observed in the presence of 1 equiv. of K+ or Rb+ were not significantly different from that without any guest, supporting the reaction first mechanism (Fig. 10A).44 Thus, the mechanism of guest binding can be switched simply by changing the guest ions.

In enzymatic reactions, substrate binding is often accompanied by structural changes in the host protein. The well-known induced-fit mechanism describes a structural change in a protein that occurs as a result of substrate binding. More recently, an alternative mechanism known as conformational selection has been proposed, in which the conformational change precedes substrate binding.46 The reaction first and recognition first mechanisms proposed in this study correspond to the conformational selection and induced-fit mechanisms, respectively, in enzymatic systems. Notably, this study provides the first demonstration that switching between these two mechanisms can be achieved simply by selecting different guest ions.

2.5. Ligand exchange of bis(saloph) cobalt(III) metallohosts with bridging diamines for guest recognition control

As described in the previous sections, L3Co2(pip)4 proved useful as the starting compound for the post-metalation modification to introduce various types of monodentate ligands, particularly primary amines. Its high reactivity toward primary amines also enabled the incorporation of diamine molecules bearing two terminal primary amino groups. Indeed, ligand exchange of L3Co2(pip)4 with 1,6-hexanediamine (DA1) afforded a doubly bridged metallohost, L3Co2(DA1)2, in 77% yield (Fig. 11a). This product has two bridging DA1 ligands on both faces of the macrocyclic plane.47
image file: d5dt01501d-f11.tif
Fig. 11 Formation of doubly and singly bridged metallohosts with the 1,6-hexanediamine ligand (DA1) based on the L3Co2 macrocycle.

The utility of the ligand exchange strategy for synthesizing the doubly bridged complex L3Co2(DA1)2 was demonstrated by a failed attempt to obtain it directly from its components, H4L3, Co(OAc)2, and DA1, under aerobic conditions (Fig. 11b). This reaction instead yielded a singly bridged species, formulated as L3Co2H(DA1)2(OAc) (Fig. 11c). These results suggested that the singly bridged species is thermodynamically more favored in the presence of AcO than the doubly bridged species. Indeed, the doubly bridged species L3Co2(DA1)2 was gradually converted into the singly bridged species L3Co2(DA1)2(OAc) in the presence of AcO, although this conversion was slow and incomplete (28% conversion after 24 h) (Fig. 11d).

Among the two types of bridged metallohosts, the doubly bridged species, L3Co2(DA1)2, showed a lower binding affinity for Na+ (Ka = 360 M−1), which may be attributed to hindered access of counter anions due to the bridging ligands DA1. The Na+ binding affinity of the singly bridged species, L3Co2H(DA1)2(OAc), was only slightly higher (Ka = 670 M−1), which can be rationalized by electrostatic repulsion from the protonated amino group (NH3+) at the terminus of the non-bridging diamine ligand DA1. Indeed, once this complex was deprotonated with Bu4NOAc (Fig. 11e), the resulting species, L3Co2(DA1)2(OAc), showed a significantly higher binding affinity for Na+ (Ka = 64[thin space (1/6-em)]000 M−1). Thus, the gate-opening from the doubly bridged to the singly bridged structure significantly enhanced guest binding affinity.47

As described in the previous sections, the guest binding affinity and ligand exchange reactivity enhanced each other in the case of the piperidine-coordinating complex, L3Co2(pip)4. A similar mutual enhancement was observed between the gate-opening reactivity of the doubly bridged species, L3Co2(DA1)2, and the Na+ binding in the cavity. While the conversion of the doubly bridged species L3Co2(DA1)2 into the singly bridged species L3Co2(DA1)2(OAc) was slow (28% conversion after 24 h; k = 3.7 × 10−6 s−1) as mentioned above (Fig. 11d), this reaction was significantly accelerated in the presence of Na+, by approximately 75-fold (k = 2.8 × 10−4 s−1). Thus, the Na+ binding and the closed → open conversion of the doubly bridged species L3Co2(DA1)2 proceeded in a mutually promoted manner.47

2.6. Shape-complementary introduction of bridging ligands into bis(saloph) cobalt(III) macrocycles

A larger class of macrocyclic bis(saloph) ligands was also employed in the synthesis of dinuclear structures, in which the two bridging ligands connect the cobalt(III) ions.48 This doubly bridged complex was synthesized by the reaction of macrocycle H4L4 (Fig. 2) with Co(OAc)2 followed by the reaction with the bis-pyridine ligand BP1 (Fig. 12a). The resulting complex, L4aCo2(BP1)2 (Fig. 12b), was characterized by NMR spectroscopy and mass spectrometry as well as X-ray crystallography. Given its structural similarity to Pd2L4 lantern-shaped cage structures,49 the doubly bridged dinuclear cobalt(III) complex L4aCo2(BP1)2 can be regarded as a trans-A2B2-type heteroleptic cage, because two types of arms alternately connect the two cobalt(III) ions. Previously, special strategies such as shape-complementary assembly were required for the synthesis of such heteroleptic cages.50
image file: d5dt01501d-f12.tif
Fig. 12 Lantern-shaped cage structures based on the bis(saloph) macrocycle H4L4. (a) Bis(pyridine) ligands used as bridging ligands. (b) Symmetric doubly bridged complex L4aCo2(BP1)2. (c) Unsymmetrical doubly bridged complexes L4Co2(BP2)(BP3), which can be regarded as a lantern-shaped cage incorporating three or four different arms.

Whereas the L4aCo2(BP1)2 complex has two identical bridging BP1 ligands on both faces, two different bridging ligands can also be selectively introduced into the L4aCo2 macrocycle via the shape-complementary assembly approach.51 Indeed, a heteroleptic doubly bridged complex, L4aCo2(BP2)(BP3), was successfully synthesized (Fig. 12c) using a similar protocol with a combination of the convergent-shaped BP2, which is based on carbazole, and the divergent-shaped BP3, which is based on fluorene (Fig. 12a).48 X-ray crystallography clearly demonstrated the face-selective bridging of the two different bis-pyridine ligands, each geometrically suited to the bent L4aCo2 macrocyclic framework; the divergent-shaped BP3 binds to the concave face, while the convergent-shaped BP2 occupies the convex face. Notably, the corresponding reactions using only BP2 or BP3 with the L4aCo2 macrocycle failed to yield the doubly bridged structure. Detailed investigation revealed that the selective formation of the mixed-ligand species, L4aCo2(BP2)(BP3), is ascribed to its thermodynamic stability.

Furthermore, the unsymmetrical macrocycle H4L4b was employed in the synthesis of the mixed-ligand species L4bCo2(BP2)(BP3) (Fig. 12c). This complex has four different arms between the two cobalt(III) ions, which can be regarded as an M2ABCD-type lantern-shaped cage, a structure that was previously considered difficult to synthesize.48

3. Functioinalization of tris(saloph) cobalt(III) cages by axial coordination

3.1. Dynamic chirality inversion of helical tris(saloph) cobalt(III) cryptands by ligand exchange

A cryptand-like tris(saloph) cage ligand, H6L5 (Fig. 2),52 features a bicyclic structure in which the three arms are doubly connected through two propeller-shaped triphenylbenzene subunits. Each arm contains a tetradentate H2saloph chelate coordination site, which can accommodate a transition metal ion. The triply metalated species, L5M3, is also expected to show strong binding affinity toward cationic guest species, due to the negatively polarized phenoxo groups in the [M(saloph)] substructures, as observed in the macrocyclic bis(saloph) analogues L3M2. Indeed, the trinuclear nickel(II) complex L5Ni3 (Fig. 13a) was found to exhibit unique binding behavior toward alkali metal ions, silver ion, guanidinium ion, etc.52–54
image file: d5dt01501d-f13.tif
Fig. 13 Chemical structures of the metallocryptands, (a) L5Ni3 and (b) L5Co3A6.

This cryptand ligand H6L5 can be converted into a trinuclear cobalt(III) complex, L5Co3A6. The six axial ligands A, coordinating to the three cobalt(III) ions in the saloph arms (Fig. 13b), can be replaced or modified, enabling functionalization and structural conversion of the cryptand framework. In fact, the trinuclear complexes L5Co3A6 (A = Me2NH, pip) were synthesized by the reaction of H6L5 with cobalt(II) acetate in the presence of appropriate amines A under aerobic conditions.55

The three [Co(saloph)A2]+ arms of the trinuclear complexes L5Co3A6 form a triple helix that extends toward the two propeller-shaped triphenylbenzene cores at the bridgeheads of the cryptand. In fact, the triple helical structure of the corresponding nickel(II) analogue, L5Ni3, was confirmed by X-ray crystallography. This structure exhibits dynamic behavior, enabling interconversion between the P and M forms.52,53 The present cobalt(III) analogues L5Co3A6, which contain achiral amine ligands (A = Me2NH, pip), also adopt a similar helical structure. The P and M forms constitute an enantiomeric pair, yielding an equilibrated racemic mixture due to the dynamic chirality interconversion. However, ligand exchange of the [Co(saloph)A2]+ (A = Me2NH, pip) arms with a chiral amine ligand makes the P and M forms diastereomeric, thereby shifting the P/M equilibrium accordingly (Fig. 14a).


image file: d5dt01501d-f14.tif
Fig. 14 Control of the P/M ratio of the helical metallocryptand L5Co3A6 via ligand exchange at the cobalt centers.

Indeed, the addition of chiral amines, S-A1 or S-A2, to L5Co3A6 (A = Me2NH, pip) resulted in ligand exchange with these chiral amines, accompanied by a gradual P/M equilibrium shift (Fig. 14a).55 The progress of ligand exchange at the cobalt(III) centers was easily monitored by 1H NMR spectroscopy as well as mass spectrometry. For example, the dimethylamine-coordinating complex L5Co3(Me2NH)6 was gradually converted to L5Co3(S-A2)6 upon addition of 12 equiv. of the chiral amine S-A2. This ligand exchange induced an equilibrium shift toward the P form, as evidenced by the growth of the negative CD signal at 550 nm (Fig. 15a(ii)). Time-course measurements indicated that the ligand exchange was almost completed within 3 h.


image file: d5dt01501d-f15.tif
Fig. 15 Time-dependent changes in CD intensity at 550 nm observed during the ligand exchange of L5Co3A6. (a) CD intensity of L5Co3A6 (A = Me2NH, pip) plotted versus time after the addition of S-A1 or S-A2 (12 equiv.). (b) CD intensity of L5Co3(Me2NH)6 plotted versus time after sequential addition of S-A1 followed by a second chiral amine R-A1 or R-A2 (120 equiv.). Adapted with permission from ref. 55.

One notable advantage of this system is that the rate of the P/M equilibrium shift can be modulated by selecting different combinations of the initial achiral amine A and the added chiral amine (Fig. 15a(i–iv)). For example, when the chiral amine S-A1 was added instead of S-A2 to L5Co3(Me2NH)6, the equilibrium shift proceeded more rapidly. In contrast, when the piperidine-coordinating complex L5Co3(pip)6 was used instead of L5Co3(Me2NH)6 as the initial complex, the reaction became significantly slower. As a result, among the four possible combinations, the rate of the P/M equilibrium shift differed by up to 60-fold.55

The ligand exchange strategy was also effective in inducing P/M chirality inversion when chiral amines with the opposite stereoconfigurations were used as the initial and added chiral sources (Fig. 14b). In fact, a P-major mixture was first prepared by the addition of the chiral amine S-A1 to racemic L5Co3(Me2NH)6, which was then inverted to an M-favored mixture upon addition of a large excess of another chiral amine, R-A1, possessing the opposite stereoconfiguration. This P/M chirality inversion was monitored by time-dependent CD spectroscopy (Fig. 15b); upon addition of the second amine, the negative CD signal at 550 nm decreased and inverted within 8 min, then became nearly constant after 1 h. When another second chiral amine, R-A2, was used instead of R-A1, the PM chirality inversion occurred 6 times more slowly. Thus, the L5Co3A6 complex was shown to be a useful helical molecular platform56,57 capable of modulating the response speeds of both the P/M equilibrium shifts and the chirality inversion through the choice of initial and added amines.55

The reverse ligand exchange, i.e., replacing a chiral ligand with an achiral ligand, showed a unique time-dependent change. When the achiral amine, piperidine, was added to L5Co3(S-A1)6, the six chiral amine ligands (S-A1) were gradually replaced by piperidine, yielding L5Co3(pip)6, which no longer contained any chiral sources in its molecular structure. This process converted a diastereomeric pair of L5Co3(S-A1)6 with a biased P/M ratio (P/M = 88[thin space (1/6-em)]:[thin space (1/6-em)]12) into an enantiomeric pair, which, in principle, could afford a racemic P/M mixture (Fig. 14c), assuming rapid P/M interconversion. Thus, the CD signal was expected to monotonically decrease and eventually become silent along the progress of the ligand exchange. However, the CD intensity exhibited an unusual irregular time-dependent change beyond expectations.58

As already stated, L5Co3(S-A1)6 exhibited a negative CD signal at 550 nm due to its P-preference (Fig. 16a(i)), but the addition of a large excess of piperidine as an achiral amine (120 equiv.) caused an immediate decrease in the CD signal followed by a reversal instead of a simple monotonic decay to zero. The resulting positive signal reached its maximum intensity after 2 h (Fig. 16a(ii)), and then gradually diminished, eventually becoming CD-silent after 3 d (Fig. 16b(i, ii) and c). This behavior indicated that the predominant chirality transiently shifted from P to M before complete racemization occurred (Fig. 17).58 Normally, when an optically active chiral compound undergoes racemization, its optical purity monotonically decreases to zero, and the chirality never inverts during the racemization process. Such a sign inversion before reaching equilibrium position is commonly observed in physical phenomena, such as damped oscillations, but is quite rare in the relaxation behavior of chemical reactions.


image file: d5dt01501d-f16.tif
Fig. 16 Time-dependent changes in the CD spectra of metallocryptand L5Co3(S-A1)6 after addition of piperidine (120 equiv.). (a) Full CD spectra for 0–120 min and (b) after 120 min. (c) CD intensity at 550 nm plotted versus time. Adapted with permission from ref. 58.

image file: d5dt01501d-f17.tif
Fig. 17 Overview of the transient P/M chirality inversion observed during the racemization of L5Co3(S-A1)6 via ligand exchange with piperidine.

This unusual and irregular time-course change during the racemization of L5Co3(S-A1)6 was investigated in detail by spectroscopic techniques, which clearly demonstrated that the six-step ligand exchange with piperidine was completed within 1 h to afford L5Co3(pip)6, i.e., at an early stage of the entire time-dependent process. The rate constants for all the ligand exchange steps were determined, revealing that the tetra-exchanged intermediate, L5Co3(S-A1)2(pip)4, which still retained two chiral S-A1 ligands, preferred the reversed stereoconfiguration, namely the M helix (Fig. 17a). It then lost all the chiral S-A1 ligands while maintaining this M-biased configuration to afford L5Co3(pip)6 (Fig. 17b), which could undergo only slow racemization over 2–3 d (Fig. 17c) due to the steric hindrance imposed by the six bulky piperidine ligands.58

Considering the above findings, opposite chiralities emerged during the forward (achiral → chiral; Fig. 14a) and reverse (chiral → achiral; Fig. 14c) ligand exchange reactions between piperidine and the chiral amine S-A1. During the ligand exchange of L5Co3(pip)6 with S-A1, the P form was always dominant (Fig. 14a), resulting in a monotonic increase in the P/M ratio. In contrast, the M form transiently became dominant during the ligand exchange of L5Co3(S-A1)6 with piperidine (Fig. 14c), leading to an unexpected transient chirality inversion (Fig. 17). This behavior constitutes a hysteretic cycle, because the M form appeared only in the reverse reaction. Such an unusual time-dependent change was discovered primarily because both the multi-step ligand exchange at the cobalt(III) centers and the P/M chirality inversion of the L5Co3A6 framework occurred on a similar timescale of minutes to hours, readily observable on a human time scale.

3.2. Tris(saloph) cobalt(III) closed cages for guest uptake/release control

Closed-cage type host molecules were anticipated to be formed when diamine ligands were introduced into the cryptand cage L5Co3 in a bridging fashion between neighboring cobalt(III) ions at the cage apertures, as demonstrated in the macrocyclic dinuclear complex L3Co2(DA1)2. The resulting closed cage species, L5Co3(diamine)3 (Fig. 18), could serve as a molecular container, because guest uptake/release is completely blocked by the bridging diamine ligands, functioning much like a container with a cap or a lid. This effective gating arises from the kinetic inertness of the cobalt(III) centers, whose slow Co–N bond cleavage/formation even suppresses transient formation of open intermediates that could allow guest entry/exit. In contrast, the corresponding monoamine-coordinating analogue, L5Co3A6 (Fig. 13b), would behave as an open container, which allows rapid guest entry/exit. Thus, interconversion between the open and closed forms functions analogously to opening and closing a container lid,59,60 providing control over the uptake and release of molecular and ionic guests.
image file: d5dt01501d-f18.tif
Fig. 18 Chemical structures of closed metallocryptands, L5Co3(diamine)3.

The triply bridged closed metallohost, L5Co3(DA1)3, was synthesized by the reaction of the cryptand ligand H6L5 with cobalt(II) acetate in the presence of diamine DA1 under aerobic conditions.61 This complex was also accessible via ligand exchange of the methylamine-coordinating open-cage complex, L5Co3(MeNH2)6, with the diamine DA1. X-ray crystallography clearly revealed the triply bridged structure of this L5Co3(DA1)3, in which three DA1 molecules connect neighboring cobalt(III) ions to form a cyclic framework. These three diamine molecules are well accommodated in the grooves of the L5Co3 triple helix, nearly completely sealing the apertures (Fig. 19a). Nevertheless, sufficient space remains at the center of the L5Co3(DA1)3 cryptand to accommodate a guest species.


image file: d5dt01501d-f19.tif
Fig. 19 (a) X-ray crystal structure of L5Co3(DA1)3. (b) Guest uptake kinetics of L5Co3(DA1)3 in CD3OD. Adapted with permission from ref. 61.

As expected, the bridging ligands effectively blocked the uptake of guest species such as alkali metal ions into the cavity of L5Co3(DA1)3, as evidenced by the 1H NMR spectra recorded immediately after guest addition. Notably, larger alkali metal ions such as Cs+ and Rb+ were taken up only very slowly, requiring approximately 5 d to reach equilibrium (Fig. 19b). This indicates that the guest uptake was kinetically suppressed, rather than thermodynamically unfavorable, and that the closed cage L5Co3(DA1)3 retained its intrinsic affinity for guest binding. Kinetic analysis revealed that guest uptake into the closed cage L5Co3(DA1)3 proceeded at least 2000 times more slowly than into the corresponding open cage L5Co3(MeNH2)6 (Fig. 20).61


image file: d5dt01501d-f20.tif
Fig. 20 Slow uptake of Cs+ by closed metallocryptand L5Co3(DA1)3.

A related closed-cage metallocryptand, L5Co3(DA7)3, bearing cystamine bridging ligands DA7, was also synthesized (Fig. 18).62 This DA7 ligand contains a disulfide bond, which is known to behave dynamically in the presence of nucleophiles63 and to remain static in their absence. Indeed, the presence of a thiolate anion was found to accelerate Cs+ uptake into L5Co3(DA7)3 by a factor of 14, whereas in its absence, the bridging DA7 ligands effectively blocked Cs+ uptake, as observed for the hexanediamine analogue, L5Co3(DA1)3.

The presence of the thiolate anion promoted the disulfide exchange equilibrium, transiently generating an open-cage intermediate that allows rapid guest entry/exit through the apertures (Fig. 21). In this system, the apertures of the L5Co3 cage were initially sealed by the introduction of bridging ligands through coordination bond formation, thereby suppressing guest uptake/release. These apertures were then transiently opened in the presence of a nucleophile via disulfide exchange reactions.


image file: d5dt01501d-f21.tif
Fig. 21 Proposed mechanism for guest uptake by L5Co3(DA7)3 via disulfide exchange accelerated by a thiolate ion.

4. Conclusions

Various types of cobalt(III) complexes bearing bis- and tris(saloph) ligands have been synthesized and shown to exhibit dynamic behaviors, despite the typically inert low-spin d6 electron configuration of cobalt(III) centers. This dynamic character arises primarily from the selective exchangeability of the two axial monodentate ligands in the [Co(saloph)X2]+ units with ligands such as primary amines, pyridine, and anionic species. Such ligand exchange enables the functionalization and dynamic switching of multi-metallic structures, including reversible redox-driven dissociation/reassociation, modulation of host–guest binding affinity and kinetics, chirality control with time-dependent inversion, and gating of guest uptake through cage closure.

In some systems, guest binding also influences ligand exchange kinetics, expanding the functional scope of these coordination platforms. Thus, the controlled introduction, removal, and exchange of axial ligands has proven to be a powerful strategy for creating functional and responsive metallohosts and metallo-supramolecular systems.

At the same time, several challenges and limitations remain to be addressed. For example, the long-term stability of axial ligands under various conditions, the potential fatigue of the systems under repeated switching cycles, and limitations in guest selectivity are all important issues that require further investigation. Understanding and overcoming these challenges will be essential for translating these dynamic systems into practical stimuli-responsive materials and molecular devices.

Conflicts of interest

There are no conflicts to declare.

Data availability

Data sharing is not applicable to this article as no new data were created or analyzed in this study.

Acknowledgements

The author warmly thanks all coworkers, collaborators, and past and present group members for their contributions to the work carried out in his laboratory. Special thanks are due to Prof. Yoko Sakata (Nagoya University), Dr Naoki Ousaka (Kyushu University), and Dr Sk Asif Ikbal (Kanazawa University) for their dedicated contributions, and to Prof. Tatsuya Nabeshima (University of Tsukuba) for valuable advice and discussions. The research cited in this article was supported by the Japan Society for the Promotion of Science (KAKENHI Grants JP21H05477 and JP23H04021 [Condensed Conjugation] as well as JP16H06510 [Coordination Asymmetry], JP18H03913, JP20K21206, JP23H01972, and JP23K17928), the CHOZEN Project at Kanazawa University, and the World Premier International Research Center Initiative (WPI), MEXT, Japan.

References

  1. (a) P. Pfeiffer, E. Breith, E. Lübbe and T. Tsumaki, Liebigs Ann. Chem., 1933, 503, 84–130 CrossRef CAS; (b) P. Pfeiffer, T. Hesse, H. Pfitzner, W. Scholl and H. Thielert, J. Prakt. Chem., 1937, 149, 217–296 CrossRef CAS; (c) C. S. Marvel, S. A. Aspey and E. A. Dudley, J. Am. Chem. Soc., 1956, 78, 4905–4909 CrossRef CAS.
  2. (a) E. N. Jacobsen, Asymmetric Catalytic Epoxidation of Unfunctionalized Olefins in Catalytic Asymmetric Synthesis, ed. I. Ojima, VCH, New York, 1993, ch. 4.2, pp. 159–202 Search PubMed; (b) T. Katsuki, Coord. Chem. Rev., 1995, 140, 189–214 CrossRef CAS; (c) E. N. Jacobsen, Acc. Chem. Res., 2000, 33, 421–431 CrossRef CAS PubMed.
  3. N. Hoshino, Coord. Chem. Rev., 1998, 174, 77–108 CrossRef CAS.
  4. (a) G. Consiglio, I. P. Oliveri, S. Failla and S. di Bella, Molecules, 2019, 24, 2514 CrossRef CAS PubMed; (b) S. di Bella, Dalton Trans., 2021, 50, 6050–6063 RSC.
  5. H. Miyasaka, A. Saitoh and S. Abe, Coord. Chem. Rev., 2007, 251, 2622–2664 CrossRef CAS.
  6. (a) X. Yang, R. A. Jones and S. Huang, Coord. Chem. Rev., 2014, 273–274, 63–75 CrossRef CAS; (b) S. di Bella and I. Fragalà, Synth. Met., 2000, 115, 191–196 CrossRef CAS; (c) P. G. Lacroix, Eur. J. Inorg. Chem., 2001, 339–348 CrossRef CAS.
  7. (a) H.-Y. Yin, J. Tang and J.-L. Zhang, Eur. J. Inorg. Chem., 2017, 5085–5093 CrossRef CAS; (b) A. Erxleben, Inorg. Chim. Acta, 2018, 472, 40–57 CrossRef CAS; (c) L. A. Alfonso-Herrera, D. Hernández-Romero, J. A. Cruz-Navarro, Á. Ramos-Ligonio, A. López-Monteon, J. M. Rivera-Villanueva, D. Morales-Morales and R. Colorado-Peralta, Coord. Chem. Rev., 2024, 505, 215698 CrossRef CAS.
  8. (a) M. J. MacLachlan, Pure Appl. Chem., 2006, 78, 873–888 CAS; (b) A. W. Kleij, Chem. – Eur. J., 2008, 14, 10520–10529 CrossRef CAS PubMed; (c) A. W. Kleij, Eur. J. Inorg. Chem., 2009, 193–205 CrossRef CAS; (d) T. Glaser, Chem. Commun., 2011, 47, 116–130 RSC; (e) C. J. Whiteoak, G. Salassa and A. W. Kleij, Chem. Soc. Rev., 2012, 41, 622–631 RSC.
  9. (a) T. Nabeshima and S. Akine, Chem. Rec., 2008, 8, 240–251 CrossRef CAS PubMed; (b) S. Akine and T. Nabeshima, Dalton Trans., 2009, 10395–10408 RSC; (c) T. Nabeshima, S. Akine, C. Ikeda and M. Yamamura, Chem. Lett., 2010, 39, 10–16 CrossRef CAS; (d) S. Akine, J. Inclusion Phenom. Macrocyclic Chem., 2012, 72, 25–54 CrossRef CAS; (e) S. Akine, Metal Complexes with Oligo(salen)-Type Ligands in The Chemistry of Metal Phenolates, ed. J. Zabicky, Patai's Chemistry of Functional Groups, John Wiley & Sons Ltd, Chichester, 2018, ch. 4, vol. 2, pp. 153–194 Search PubMed; (f) M. T. Chaudhry, S. Akine and M. J. MacLachlan, Chem. Soc. Rev., 2021, 50, 10713–10732 RSC.
  10. (a) S. J. Wezenberg and A. W. Kleij, Angew. Chem., Int. Ed., 2008, 47, 2354–2364 CrossRef CAS PubMed; (b) S. M. Elbert and M. Mastalerz, Org. Mater., 2020, 2, 182–203 CrossRef CAS.
  11. (a) R. M. Haak, S. J. Wezenberg and A. W. Kleij, Chem. Commun., 2010, 46, 2713–2723 RSC; (b) A. Decortes, A. M. Castilla and A. W. Kleij, Angew. Chem., Int. Ed., 2010, 49, 9822–9837 CrossRef CAS PubMed; (c) D. E. White, P. M. Tadross, Z. Lu and E. N. Jacobsen, Tetrahedron, 2014, 70, 4165–4180 CrossRef CAS PubMed.
  12. For acyclic salen/saloph oligomers, see: (a) J. H. Chong, S. J. Ardakani, K. J. Smith and M. J. MacLachlan, Chem. – Eur. J., 2009, 15, 11824–11828 CrossRef CAS PubMed; (b) S. J. Wezenberg, G. Salassa, E. C. Escudero-Adán, J. Benet-Buchholz and A. W. Kleij, Angew. Chem., Int. Ed., 2011, 50, 713–716 CrossRef CAS PubMed; (c) T. J. Dunn, C. F. Ramogida, C. Simmonds, A. Paterson, E. W. Y. Wong, L. Chiang, Y. Shimazaki and T. Storr, Inorg. Chem., 2011, 50, 6746–6755 CrossRef CAS PubMed; (d) M. Mastalerz and I. M. Oppel, Eur. J. Org. Chem., 2011, 5971–5980 CrossRef CAS; (e) T. J. Dunn, L. Chiang, C. F. Ramogida, M. I. Webb, D. Savard, M. Sakaguchi, T. Ogura, Y. Shimazaki and T. Storr, Dalton Trans., 2012, 41, 7905–7914 RSC; (f) H. Achira, M. Ito, T. Mutai, I. Yoshikawa, K. Araki and H. Houjou, Dalton Trans., 2014, 43, 5899–5907 Search PubMed; (g) S. A. Ikbal, Y. Sakata and S. Akine, Dalton Trans., 2021, 50, 4119–4123 RSC.
  13. (a) M. E. Briggs and A. I. Cooper, Chem. Mater., 2017, 29, 149–157 CrossRef CAS PubMed; (b) M. Mastalerz, Acc. Chem. Res., 2018, 51, 2411–2422 CrossRef CAS PubMed.
  14. (a) S. J. Rowan, S. J. Cantrill, G. R. L. Cousins, J. K. M. Sanders and J. F. Stoddart, Angew. Chem., Int. Ed., 2002, 41, 898–952 CrossRef; (b) M. Mastalerz, Angew. Chem., Int. Ed., 2010, 49, 5042–5053 CrossRef CAS PubMed; (c) Y. Jin, C. Yu, R. J. Denman and W. Zhang, Chem. Soc. Rev., 2013, 42, 6634–6654 RSC.
  15. (a) A. G. Manfredotti and C. Guastini, Acta Crystallogr., Sect. C: Cryst. Struct. Commun., 1983, C39, 863–865 CrossRef CAS; (b) A. Radha, M. Seshasayee, K. Ramalingam and G. Aravamudan, Acta Crystallogr., Sect. C: Cryst. Struct. Commun., 1985, 41, 1169–1171 CrossRef; (c) J. Wang, F.-L. Bei, X.-Y. Xu, X.-J. Yang and X. Wang, J. Chem. Crystallogr., 2003, 33, 845–849 CrossRef CAS.
  16. (a) G. D. Fallon and B. M. Gatehouse, Acta Crystallogr., Sect. B, 1976, 32, 2591–2597 CrossRef; (b) N. Kumari, R. Prajapati and L. Mishra, Polyhedron, 2008, 27, 241–248 CrossRef CAS; (c) S. Mahato, P. Rawal, A. K. Devadkar, M. Joshi, A. R. Choudhury, B. Biswas, P. Gupta and T. K. Panda, Org. Biomol. Chem., 2022, 20, 1103–1111 RSC; (d) Y. Sekine, S. Kusumoto, A. Sugimoto, M. Nakaya and S. Hayami, Cryst. Growth Des., 2023, 23, 2013–2017 CrossRef CAS.
  17. (a) D. Hall and F. H. Moore, J. Chem. Soc. A, 1966, 1822–1824 RSC; (b) C. A. Barboza, J. C. Germino, A. M. Santana, F. J. Quites, P. A. M. Vazquez and T. D. Z. Atvars, J. Phys. Chem. C, 2015, 119, 6152–6163 CrossRef CAS.
  18. (a) J. Reglinski, S. Morris and D. E. Stevenson, Polyhedron, 2002, 21, 2175–2182 CrossRef CAS; (b) J. S. Matalobos, A. M. García-Deibe, M. Fondo, D. Navarro and M. R. Bermejo, Inorg. Chem. Commun., 2004, 7, 311–314 CrossRef; (c) A. W. Kleij, M. Kuil, M. Lutz, D. M. Tooke, A. L. Spek, P. C. J. Kamer, P. W. N. M. van Leeuwen and J. N. H. Reek, Inorg. Chim. Acta, 2006, 359, 1807–1814 CrossRef CAS; (d) M. Odoko, N. Tsuchida and N. Okabe, Acta Crystallogr., Sect. E: Struct. Rep. Online, 2006, E62, m708–m709 CrossRef.
  19. For a review, see: A. W. Kleij, Dalton Trans., 2009, 4635–4639 RSC.
  20. (a) A. W. Kleij, M. Kuil, D. M. Tooke, M. Lutz, A. L. Spek and J. N. H. Reek, Chem. – Eur. J., 2005, 11, 4743–4750 CrossRef CAS PubMed; (b) M. Kuil, P. E. Goudriaan, P. W. N. M. van Leeuwen and J. N. H. Reek, Chem. Commun., 2006, 4679–4681 RSC; (c) M. Kuil, I. M. Puijk, A. W. Kleij, D. M. Tooke, A. L. Spek and J. N. H. Reek, Chem. – Asian J., 2009, 4, 50–57 CrossRef CAS PubMed; (d) M. Kuil, P. E. Goudriaan, A. W. Kleij, D. M. Tooke, A. L. Spek, P. W. N. M. van Leeuwen and J. N. H. Reek, Dalton Trans., 2007, 2311–2320 RSC; (e) T. Gadzikwa, R. Bellini, H. L. Dekker and J. N. H. Reek, J. Am. Chem. Soc., 2012, 134, 2860–2863 CrossRef CAS PubMed.
  21. (a) A. W. Kleij, M. Kuil, D. M. Tooke, A. L. Spek and J. N. H. Reek, Inorg. Chem., 2007, 46, 5829–5831 CrossRef CAS PubMed; (b) J. Flapper and J. N. H. Reek, Angew. Chem., Int. Ed., 2007, 46, 8590–8592 CrossRef CAS PubMed; (c) G. Salassa, A. M. Castilla and A. W. Kleij, Dalton Trans., 2011, 40, 5236–5243 RSC; (d) I. Jacobs, A. C. T. van Duin, A. W. Kleij, M. Kuil, D. M. Tooke, A. L. Spek and J. N. H. Reek, Catal. Sci. Technol., 2013, 3, 1955–1963 RSC; (e) D. Anselmo, R. Gramage-Doria, T. Besset, M. V. Escárcega-Bobadilla, G. Salassa, E. C. Escudero-Adán, M. M. Belmonte, E. Martin, J. N. H. Reek and A. W. Kleij, Dalton Trans., 2013, 42, 7595–7603 RSC.
  22. M. Karmakar and S. Chattopadhyay, J. Mol. Struct., 2019, 1186, 155–186 CrossRef CAS.
  23. (a) M. Amirnasr, K. J. Schenk, A. Gorji and R. Vafazadeh, Polyhedron, 2001, 20, 695–702 CrossRef CAS; (b) M. Salehi, S. Meghdadi, M. Amirnasr and K. Mereiter, Acta Crystallogr., Sect. E:Struct. Rep. Online, 2009, 65, m942–m943 CrossRef CAS PubMed.
  24. Ruthenium complex was also used as a platform of functionalization, see: K. Chichak, U. Jacquemard and N. R. Branda, Eur. J. Inorg. Chem., 2002, 357–368 CrossRef CAS.
  25. (a) S. C. F. Au-Yeung and D. R. Eaton, Inorg. Chem., 1984, 23, 1517–1520 CrossRef CAS; (b) M. F. Summers, L. G. Marzilli, N. Bresciani-Pahor and L. Randaccio, J. Am. Chem. Soc., 1984, 106, 4478–4485 CrossRef CAS; (c) L. M. Manus, R. J. Holbrook, T. A. Atesin, M. C. Heffern, A. S. Harney, A. L. Eckermann and T. J. Meade, Inorg. Chem., 2013, 52, 1069–1076 CrossRef CAS PubMed.
  26. S. Akine, Dalton Trans., 2021, 50, 4429–4444 RSC.
  27. (a) E. Eichhorn, A. Rieker, B. Speiser and H. Stahl, Inorg. Chem., 1997, 36, 3307–3317 CrossRef CAS PubMed; (b) A. A. Khandar, B. Shaabani, F. Belaj and A. Bakhtiari, Inorg. Chim. Acta, 2007, 360, 3255–3264 CrossRef CAS.
  28. (a) M. J. Hardie and C. L. Raston, Angew. Chem., Int. Ed., 2000, 39, 3835–3839 CrossRef CAS; (b) D. A. Leigh, P. J. Lusby, R. T. McBurney, A. Morelli, A. M. Z. Slawin, A. R. Thomson and D. B. Walker, J. Am. Chem. Soc., 2009, 131, 3762–3771 CrossRef CAS PubMed; (c) T. Nishino, Y. Yamada, S. Akine, K. Sugimoto and K. Tanaka, Dalton Trans., 2016, 45, 3831–3837 RSC; (d) Y.-L. Lu, K. Wu, Y.-H. Huang, W.-C. Li, Z.-M. Cao, X.-H. Yan, X.-D. Zhang, C.-H. Liu, J. Ruan, H.-S. Xu, M. Pan and C.-Y. Su, J. Am. Chem. Soc., 2024, 146, 20414–20424 CrossRef CAS PubMed.
  29. H. Shimakoshi, H. Takemoto, I. Aritome and Y. Hisaeda, Tetrahedron Lett., 2002, 43, 4809–4812 CrossRef CAS.
  30. H. Shimakoshi, T. Takemoto, I. Aritome and Y. Hisaeda, Inorg. Chem., 2005, 44, 9134–9136 CrossRef CAS PubMed.
  31. H. Shimakoshi, S. Hirose, M. Ohba, T. Shiga, H. Ōkawa and Y. Hisaeda, Bull. Chem. Soc. Jpn., 2005, 78, 1040–1046 CrossRef CAS.
  32. H. Shimakoshi, K. Shibata and Y. Hisaeda, Inorg. Chem., 2009, 48, 1045–1052 CrossRef CAS PubMed.
  33. H. Shimakoshi and Y. Hisaeda, J. Synth. Org. Chem., Jpn., 2012, 70, 60–70 CrossRef CAS.
  34. (a) S. Akine, F. Utsuno and T. Nabeshima, Chem. Commun., 2010, 46, 1029–1031 RSC; (b) S. Akine, F. Utsuno, S. Piao, H. Orita, S. Tsuzuki and T. Nabeshima, Inorg. Chem., 2016, 55, 810–821 CrossRef CAS PubMed.
  35. (a) M. Cametti, Y. Sakata, J. Martí-Rujas and S. Akine, Inorg. Chem., 2019, 58, 14871–14875 CrossRef CAS PubMed; (b) M. T. Chaudhry, B. O. Patrick, S. Akine and M. J. MacLachlan, Org. Biomol. Chem., 2022, 20, 8259–8268 RSC; (c) S. Akine, M. Nakano, Y. Sakata and S. Yano, Chem. – Eur. J., 2024, 30, e202403071 CrossRef CAS PubMed.
  36. For related macrocyclic bis(saloph) hosts with different sizes, see: (a) F. C. J. M. van Veggel, M. Bos, S. Harkema, W. Verboom and D. N. Reinhoudt, Angew. Chem., Int. Ed. Engl., 1989, 28, 746–748 CrossRef; (b) F. C. J. M. van Veggel, M. Bos, S. Harkema, H. van de Bovenkamp, W. Verboom, J. Reedijk and D. N. Reinhoudt, J. Org. Chem., 1991, 56, 225–235 CrossRef CAS; (c) Y. Sakata, S. Kobayashi and S. Akine, Chem. Commun., 2017, 53, 6363–6366 RSC; (d) Y. Sakata, S. Kobayashi, M. Yamamoto, K. Doken, M. Kamezawa, S. Yamaki and S. Akine, Commun. Chem., 2024, 7, 166 CrossRef CAS PubMed; (e) S. Akine, Z. Varadi and T. Nabeshima, Eur. J. Inorg. Chem., 2013, 5987–5998 CrossRef CAS.
  37. For representative macrocyclic tris(saloph) hosts, see: (a) S. Akine, T. Taniguchi and T. Nabeshima, Tetrahedron Lett., 2001, 42, 8861–8864 CrossRef CAS; (b) A. J. Gallant and M. J. MacLachlan, Angew. Chem., Int. Ed., 2003, 42, 5307–5310 Search PubMed; (c) S. Akine, D. Hashimoto, T. Saiki and T. Nabeshima, Tetrahedron Lett., 2004, 45, 4225–4227 Search PubMed; (d) T. Nabeshima, H. Miyazaki, A. Iwasaki, S. Akine, T. Saiki, C. Ikeda and S. Sato, Chem. Lett., 2006, 35, 1070–1071 Search PubMed; (e) S. Akine, S. Sunaga, T. Taniguchi, H. Miyazaki and T. Nabeshima, Inorg. Chem., 2007, 46, 2959–2961 CrossRef CAS PubMed; (f) A. Yamashita, A. Watanabe, S. Akine, T. Nabeshima, M. Nakano, T. Yamamura and T. Kajiwara, Angew. Chem., Int. Ed., 2011, 50, 4016–4019 CrossRef CAS PubMed; (g) S. Akine, S. Sunaga and T. Nabeshima, Chem. – Eur. J., 2011, 17, 6853–6861 CrossRef CAS PubMed; (h) T. Nakamura, S. Yano, Y.-F. Liu and T. Nabeshima, Inorg. Chem., 2024, 63, 12697–12702 CrossRef CAS PubMed.
  38. For related macrocyclic tetrakis(saloph) hosts, see: (a) S.-S. Sun, C. L. Stern, S. T. Nguyen and J. T. Hupp, J. Am. Chem. Soc., 2004, 126, 6314–6326 CrossRef CAS PubMed; (b) S.-i. Kawano, Y. Ishida and K. Tanaka, J. Am. Chem. Soc., 2015, 137, 2295–2302 CrossRef CAS PubMed; (c) T. Nakamura, S. Tsukuda and T. Nabeshima, J. Am. Chem. Soc., 2019, 141, 6462–6467 CrossRef CAS PubMed; (d) S.-i. Kawano, M. Nakaya, M. Saitow, A. Ishiguro, T. Yanai, J. Onoe and K. Tanaka, J. Am. Chem. Soc., 2022, 144, 6749–6758 CrossRef CAS PubMed.
  39. Y. Sakata, C. Murata and S. Akine, Nat. Commun., 2017, 8, 16005 CrossRef CAS PubMed.
  40. (a) C. Marquez and W. M. Nau, Angew. Chem., Int. Ed., 2001, 40, 3155–3160 CrossRef CAS; (b) O. Danylyuk, V. P. Fedin and V. Sashuk, Chem. Commun., 2013, 49, 1859–1861 RSC.
  41. D. Walter, Y. Sakata and S. Akine, Chem. – Asian J., 2025, 20, e202401876 CrossRef CAS PubMed.
  42. (a) D. A. Roberts, B. S. Pilgrim and J. R. Nitschke, Chem. Soc. Rev., 2018, 47, 626–644 RSC; (b) M.-M. Gan, J.-Q. Liu, L. Zhang, Y.-Y. Wang, F. E. Hahn and Y.-F. Han, Chem. Rev., 2018, 118, 9587–9641 CrossRef CAS PubMed.
  43. Y. Sakata, M. Okada, M. Tamiya and S. Akine, Chem. – Eur. J., 2020, 26, 7595–7601 CrossRef CAS PubMed.
  44. Y. Sakata, M. Tamiya, M. Okada and S. Akine, J. Am. Chem. Soc., 2019, 141, 15597–15604 CrossRef CAS PubMed.
  45. (a) P. J. Tummino and R. A. Copeland, Biochemistry, 2008, 47, 5481–5492 CrossRef CAS PubMed; (b) G. G. Hammes, Y.-C. Chang and T. G. Oas, Proc. Natl. Acad. Sci. U. S. A., 2009, 106, 13737–13741 CrossRef CAS PubMed; (c) A. D. Vogt and E. di Cera, Biochemistry, 2012, 51, 5894–5902 CrossRef CAS PubMed.
  46. (a) L. C. James and D. S. Tawfik, Trends Biochem. Sci., 2003, 28, 361–368 CrossRef CAS PubMed; (b) D. D. Boehr, R. Nussinov and P. E. Wright, Nat. Chem. Biol., 2009, 5, 789–796 CrossRef CAS PubMed; (c) P. Csermely, R. Palotai and R. Nussinov, Trends Biochem. Sci., 2010, 35, 539–546 CrossRef CAS PubMed.
  47. Y. Sakata, M. Okada and S. Akine, Chem. – Eur. J., 2021, 27, 2284–2288 CrossRef CAS PubMed.
  48. B. Zhang, H. Lee, J. J. Holstein and G. H. Clever, Angew. Chem., Int. Ed., 2024, 63, e202404682 CrossRef CAS PubMed.
  49. D. A. McMorran and P. J. Steel, Angew. Chem., Int. Ed., 1998, 37, 3295–3297 CrossRef CAS PubMed.
  50. (a) W. M. Bloch and G. H. Clever, Chem. Commun., 2017, 53, 8506–8516 RSC; (b) S. Pullen, J. Tessarolo and G. H. Clever, Chem. Sci., 2021, 12, 7269–7293 RSC.
  51. (a) W. M. Bloch, Y. Abe, J. J. Holstein, C. M. Wandtke, B. Dittrich and G. H. Clever, J. Am. Chem. Soc., 2016, 138, 13750–13755 CrossRef CAS PubMed; (b) W. M. Bloch, J. J. Holstein, W. Hiller and G. H. Clever, Angew. Chem., Int. Ed., 2017, 56, 8285–8289 CrossRef CAS PubMed.
  52. (a) S. Akine, M. Miyashita, S. Piao and T. Nabeshima, Inorg. Chem. Front., 2014, 1, 53–57 RSC; (b) S. Akine, M. Miyashita and T. Nabeshima, Inorg. Chem., 2021, 60, 12961–12971 CrossRef CAS PubMed.
  53. S. A. Ikbal, P. Zhao, M. Ehara and S. Akine, Sci. Adv., 2023, 9, eadj5536 CrossRef CAS PubMed.
  54. For related oligo(salen)/oligo(saloph) cages, see: (a) S. Akine, S. Piao, M. Miyashita and T. Nabeshima, Tetrahedron Lett., 2013, 54, 6541–6544 CrossRef CAS; (b) S. M. Elbert, F. Rominger and M. Mastalerz, Chem. – Eur. J., 2014, 20, 16707–16720 CrossRef CAS PubMed; (c) J. Jiao, C. Tan, Z. Li, Y. Liu, X. Han and Y. Cui, J. Am. Chem. Soc., 2018, 140, 2251–2259 Search PubMed; (d) S. M. Elbert, W.-S. Zhang, Y. Vaynzof, N. Oberhof, M. Bernhardt, M. Pernpointner, F. Rominger, R. R. Schröder and M. Mastalerz, Chem. Mater., 2019, 31, 6210–6223 CrossRef CAS; (e) Y. Sakata, R. Tsuyuki, S. Sugimoto and S. Akine, Chem. Commun., 2021, 57, 13510–13513 RSC; (f) F. Uhrmacher, S. M. Elbert, F. Rominger and M. Mastalerz, Eur. J. Inorg. Chem., 2022, e202100864 CrossRef CAS; (g) S. Firdausiah, R. Ide, M. Ehara, P. Zhao, Y. Sakata and S. Akine, Chem. – Eur. J., 2025, 31, e202501277 CrossRef CAS PubMed.
  55. Y. Sakata, S. Chiba, M. Miyashita, T. Nabeshima and S. Akine, Chem. – Eur. J., 2019, 25, 2962–2966 CrossRef CAS PubMed.
  56. (a) H. Miyake and H. Tsukube, Chem. Soc. Rev., 2012, 41, 6977–6991 RSC; (b) S. Akine, Inorganics, 2018, 6, 80 CrossRef; (c) S. Akine and H. Miyake, Coord. Chem. Rev., 2022, 468, 214582 CrossRef CAS; (d) S. Akine, Pure Appl. Chem., 2023, 95, 331–341 CrossRef CAS.
  57. Recent examples of chirality control of metal complexes, see: (a) S. Akine, T. Taniguchi and T. Nabeshima, Tetrahedron Lett., 2006, 47, 8419–8422 CrossRef CAS; (b) S. Akine, T. Matsumoto and T. Nabeshima, Chem. Commun., 2008, 4604–4606 RSC; (c) S. Akine, S. Hotate, T. Matsumoto and T. Nabeshima, Chem. Commun., 2011, 47, 2925–2927 RSC; (d) S. Akine, S. Hotate and T. Nabeshima, J. Am. Chem. Soc., 2011, 133, 13868–13871 CrossRef CAS PubMed; (e) S. Akine, S. Sairenji, T. Taniguchi and T. Nabeshima, J. Am. Chem. Soc., 2013, 135, 12948–12951 CAS; (f) S. Sairenji, S. Akine and T. Nabeshima, Tetrahedron Lett., 2014, 55, 1987–1990 CAS; (g) S. Sairenji, S. Akine and T. Nabeshima, Chem. Lett., 2014, 43, 1107–1109 CrossRef CAS; (h) S. Akine, T. Matsumoto and T. Nabeshima, Angew. Chem., Int. Ed., 2016, 55, 960–964 CrossRef CAS PubMed; (i) S. Sairenji, S. Akine and T. Nabeshima, Dalton Trans., 2016, 45, 14902–14906 RSC; (j) S. Sairenji, S. Akine and T. Nabeshima, Sci. Rep., 2018, 8, 137 CrossRef PubMed; (k) S. Akine, K. Nomura, M. Takahashi, Y. Sakata, T. Mori, W. Nakanishi and K. Ariga, Dalton Trans., 2023, 52, 260–268 RSC; (l) S. Muratsugu, K. Sawaguchi, T. Shiraogawa, S. Chiba, Y. Sakata, S. Shirai, H. Baba, M. Ehara, S. Akine and M. Tada, Chem. Commun., 2024, 60, 2094–2097 RSC.
  58. Y. Sakata, S. Chiba and S. Akine, Proc. Natl. Acad. Sci. U. S. A., 2022, 119, e2113237119 CrossRef CAS PubMed.
  59. S. Akine and Y. Sakata, Chem. Lett., 2020, 49, 428–441 CrossRef CAS.
  60. (a) E. L. Piatnitski and K. D. Deshayes, Angew. Chem., Int. Ed., 1998, 37, 970–972 CrossRef CAS; (b) S. Zarra, M. M. J. Smulders, Q. Lefebvre, J. K. Clegg and J. R. Nitschke, Angew. Chem., Int. Ed., 2012, 51, 6882–6885 CrossRef CAS PubMed; (c) H. Wang, F. Liu, R. C. Helgeson and K. N. Houk, Angew. Chem., Int. Ed., 2013, 52, 655–659 CrossRef CAS PubMed.
  61. S. Akine, M. Miyashita and T. Nabeshima, J. Am. Chem. Soc., 2017, 139, 4631–4634 CrossRef CAS PubMed.
  62. S. Akine, M. Miyashita and T. Nabeshima, Chem. – Eur. J., 2019, 25, 1432–1435 CrossRef CAS PubMed.
  63. S. P. Black, J. K. M. Sanders and A. R. Stefankiewicz, Chem. Soc. Rev., 2014, 43, 1861–1872 RSC.

Footnotes

H2salen = N,N′-disalicylideneethylenediamine.
H2saloph = N,N′-disalicylidene-o-phenylenediamine.

This journal is © The Royal Society of Chemistry 2025
Click here to see how this site uses Cookies. View our privacy policy here.