Introduction

Macrocyclic chemistry has become a major topic of supramolecular chemistry because of the unique recognition properties [1], serving as self-assembly building blocks [2,3,4] and potential applications in biomedicine [5,6,7], and material science [8,9,10,11]. A lot of macrocycles, typically crown ethers, cyclodextrins, calixarenes, and cucurbiturils, have been developed and investigated extensively [12]. In recent years, stimuli-responsive molecular recognition and self-assembly have gained more and more attention, and demonstrated abundant applications in various fields. Conventional stimulating modalities include temperature, pH, light, electric, enzyme, and so on. Among these stimuli, light is especially appealing for its cleanness, non-invasion and high spatiotemporal resolution [13, 14]. The most commonly used photo-responsive molecules include azobenzene, stilbene, and spiropyran and diarylethene derivatives.

Azo compounds represent more than half of the world production of industrial dyes [15] and most of them displayed reversibly light-driven trans- to cis-isomerization. Cis-isomer is substantially different from trans-isomer in many aspects: larger dipole moment, weaker π–π* and stronger n–π* absorption, more bent geometry and typically less thermodynamically stable. Therefore, they have been widely used to drive photo-responsive molecular switches [16, 17], polymers [18], metal–organic materials [19], liquid crystals [20], and molecular machines [21].

Decorating azo groups on macrocycles is a convenient and powerful way to endow macrocycles with additional physical and chemical properties, such as larger cavity for efficient recognition, guest induced color change for optical sensing, and more importantly, photo-responsive switch of binding affinity, self-assembly behavior and function [22]. Several excellent reviews have summarized the progress in this field [23, 24], but no review summarized the azo-containing macrocycles comprehensively and highlighted their recognition, assembly and application. In this review, we will summarize the structures of azo-containing macrocyclic compounds reported up to now and focus our special attention on their recognition, self-assembly and application. The structure of this review will be such that we first summarize and comprehensively list the structures of different classes of azo macrocycles, including azo crown ether, azo cyclodextrin, azo calixarene, azo resorcinarene, azo calixpyrrole, azo benzenophane, azo porphyrin and others. Then we describe selectively a series of representative works on azo macrocycles with their molecular recognition, self-assembly and application.

Structures of azo macrocycles

Azo crown ethers

Crown ethers, cyclic chemical compounds containing a ring with several ether groups, are the first generation of macrocyclic compounds. In 1967, Pedersen discovered crown ethers and found their binding with alkali metal cation [25]. Crown ethers are represented as [m]crown-n, where m is the total number of atoms and n is the number of oxygen atoms. Crown ethers have been applied in phase transfer catalysis, the decontamination of nuclear waste, and so on [12] (Schemes 1, 2, 3, 4, 5, 6, 7, 8, 9; Tables 1, 2, 3, 4, 5, 6, 7, 8, 9).

Scheme 1
scheme 1

Structures of crown ethers with one azo group in a simple ring

Table 1 Azo crown ethers in Scheme 1
Scheme 2
scheme 2

Structures of crown ethers with two azo groups in a simple ring

Table 2 Azo crown ethers in Scheme 2
Scheme 3
scheme 3

Structures of bis(crown ethers) with one azo group in a simple ring

Table 3 Azo crown ethers in Scheme 3
Scheme 4
scheme 4

Structures of bis(crown ethers) with azo groups in complex ring

Table 4 Azo crown ethers in Scheme 4
Scheme 5
scheme 5

Structures of bis(crown ethers) linked by azo group

Table 5 Azo crown ethers in Scheme 5
Scheme 6
scheme 6scheme 6

Structures of crown ethers with one azo group on its periphery

Table 6 Azo crown ethers in Scheme 6
Scheme 7
scheme 7

Structures of azo crown ethers linked with polymers

Table 7 Azo crown ethers in Scheme 7
Scheme 8
scheme 8

Structures of crown ethers with two azo groups on its periphery

Table 8 Azo crown ethers in Scheme 8
Scheme 9
scheme 9

Structures of crown ethers with azo groups in complex ring

Table 9 Azo crown ethers in Scheme 9

Azo cyclodextrins

Cyclodextrins, the second generation of macrocyclic compounds, are cyclic oligosaccharides consisting of d-glucose units linked by α-(1,4)-glucose bonds. The most common cyclodextrins consist 6, 7, and 8 glucose units, and are called α-, β-, and γ-cyclodextrins, respectively. They are cylinder shaped and possess hydrophobic inner cavity and hydrophilic outer surface, resulting in their unique complex ability with a variety of lipophilic molecules [164] (Schemes 10, 11; Tables 10, 11).

Scheme 10
scheme 10scheme 10

Structures of azo cyclodextrins with one host unit

Table 10 Azo cyclodextrins in Scheme 10
Scheme 11
scheme 11

Structures of azo cyclodextrins with more than one host unit

Table 11 Azo cyclodextrins in Scheme 11

Azo calixarenes

Calixarenes are the third generation of macrocyclic compounds composed of phenolic units bridged with methylene groups at o-positions of phenolic hydroxyl groups. Calixarenes with n phenolic units are represented as calix[n]arene and the most studied calixarenes have n = 4, 5, 6, and 8. Calixarenes are facilely modified to introduce various kinds of functional groups and to adjust cavity sizes, which have been described as having “(almost) unlimited possibilities” [213]. Properly modified calixarenes can bind with both metal ions like crown ethers, and hydrophobic molecules like cyclodextrins. A unique property of calixarene is its various conformation. The smallest calix[4]arene have four typical conformations: cone, partial cone, 1,2-alterative, 1,3-alterative [214], while other larger calixarenes have more complex conformational space (Schemes 12, 13, 14, 15, 16, 17, 18; Tables 12, 13, 14, 15, 16, 17, 18).

Scheme 12
scheme 12

Structures of azo calixarenes (part 1)

Table 12 Azo calixarenes in Scheme 12
Scheme 13
scheme 13

Structures of azo calixarenes (part 2)

Table 13 Azo calixarenes in Scheme 13
Scheme 14
scheme 14

Structures of azo calixarenes (part 3)

Table 14 Azo calixarenes in Scheme 14
Scheme 15
scheme 15scheme 15

Structures of azo calixarenes (part 4)

Table 15 Azo calixarenes in Scheme 15
Scheme 16
scheme 16

Structures of azo calixarenes (part 5)

Table 16 Azo calixarenes in Scheme 16
Scheme 17
scheme 17scheme 17

Structures of azo calixarenes (part 6)

Table 17 Azo calixarenes in Scheme 17
Scheme 18
scheme 18scheme 18scheme 18scheme 18scheme 18

Structures of azo calixarenes (part 7)

Table 18 Azo calixarenes in Scheme 18

Azo resorcinarenes

Resorcinarenes are a class of macrocycles which are structurally similar to calixarenes. In contrast to calixarenes, the repeat unit of resorcinarenes is resorcinol instead of phenol. Resorcinarenes with n repeat units are represented as resorcin[n]arenes and the most studied resorcinarenes are the resorcin[4]arene [399]. In general, resorcinarenes can be synthesized by acid-catalyzed condensation of resorcinol or its derivatives with aldehyde [400]. Various methods were developed for their functionalization to synthesize sophisticated derivatives [399] (Schemes 19; Table 19).

Scheme 19
scheme 19scheme 19

Structures of azo resorcinarenes

Table 19 Azo resorcinarenes in Scheme 19

Azo calixpyrroles

Calixpyrroles are a class of hetero-calixarenes containing pyrrole units linked by meso-carbon bridges. Calixpyrroles can be obtained by the condensation of pyrrole and ketones. Sessler and co-workers first found they can bind anion guests [424] and be used to construct ion-pair receptors [425]. They have been utilized for sensing [426], extracting [427] and transporting [428] of anions (Scheme 20, Table 20).

Scheme 20
scheme 20

Structures of azo calixpyrroles

Table 20 Azo calixpyrroles in Scheme 20

Azobenzenophanes

Azobenzenophanes are cyclic oligomer of azobenzenes with unique isomerization properties [22]. They have potential applications in constructing multistate switches and overcoming the instability of Z-isomer of azobenzene (Scheme 21, Table 21).

Scheme 21
scheme 21scheme 21

Structures of azobenzenophanes

Table 21 Azobenzenophanes in Scheme 21

Azo porphyrins

Porphyrins are a class of planar macrocycle with four pyrrolic subunits bridged by methine units. Porphyrins are involved in various biological processes, such as oxygen binding, photosynthesis and electron transfer. They are also used as photosensitizer in photodynamic therapy and artificial light harvesting system (Scheme 22, Table 22).

Scheme 22
scheme 22scheme 22

Structures of azo porphyrins

Table 22 Azo porphyrins in Scheme 22

Others

Although the above compounds covered most of azo macrocycles, some azo containing macrocyclic molecules not belong to derivatives of typical macrocycles. Here, we classified these compounds as others (Schemes 23, 24, 25, 26, 27; Tables 23, 24, 25, 26, 27).

Scheme 23
scheme 23scheme 23

Structures of other azo macrocycles (part 1)

Table 23 Other azo macrocycles in Scheme 23
Scheme 24
scheme 24

Structures of other azo macrocycles (part 2)

Table 24 Other azo macrocycles in Scheme 24
Scheme 25
scheme 25

Structures of other azo macrocycles (part 3)

Table 25 Other azo macrocycles in Scheme 25
Scheme 26
scheme 26

Structures of other azo macrocycles (part 4)

Table 26 Other azo macrocycles in Scheme 26
Scheme 27
scheme 27scheme 27

Structures of other azo macrocycles (part 5)

Table 27 Other azo macrocycles in Scheme 27

Recognition, self-assembly and application

In this section, we describe selectively the following 30 representative works on azo-containing macrocycles to demonstrate their recognition, self-assembly and application.

Recognition

Early in 1979, Ueno et al. reported an azobenzene-capped β-cyclodextrin (β-CD) that can regulate the 1:2 host–guest complexation and its binding ability by cis/trans photoisomerization of azobenzene in response to light stimulus (Fig. 1) [191]. Reaction of β-CD with 4,4′-bis(chlorocarbonyl)-trans-azobenzene afforded compound 349 (see Scheme 10) in 20% yield. Upon irradiation, 349 was converted into its cis-isomer with a much larger cavity. For most of tested aromatic and olefinic guests, the cavity of cis-isomer is large enough to encapsulate two guests. All tested guests show higher binding affinities with the cis-isomer than with the trans-isomer. The most striking feature took place in binding 4,4′-bipyridine. 349 cannot bind 4,4′-bipyridine when the azobenzene group is in its trans form. However, upon irradiation with UV light, it converts to cis-isomer, and then the host includes 4,4′-bipyridine in its expanded cavity, as indicated by circular dichroism experiments. When cis 349 reverses to its trans-isomer in the dark, 4,4′-bipyridine is ejected from the cavity.

Fig. 1
figure 1

Reproduced with permission from Ref. [191]. Copyright 1979 from American Chemical Society

Illustration of isomerization and complexation with guest of 349.

In 1992, the same group synthesized a color change indicator in acidic solution 325 (see Scheme 10) by appending one methyl red dye to 6-deoxy-6-amino-β-CD (Fig. 2) [165]. Methyl red shows color change depending on the pH conditions. The dye moiety is included in the hydrophobic cavity of β-CD, which protects it from protonation. Therefore, the dye-appended β-CD remains yellow in acidic solution. The color change will occur when an organic guest displaces the methyl red dye from the interior of β-CD, which makes the dye moiety available to protonation. It should be possible to exploit the molecular recognition capability of β-CD to develop a range of such indicators.

Fig. 2
figure 2

Reproduced with permission from Ref. [165]. Copyright 1992 from Nature Publishing Group

Illustration of the guest-induced conformational change of 325 that causes the color change.

In 2017, Hayashita and co-workers reported an azobenzene bearing γ-CD derivative 350 (see Scheme 10) for recognizing phosphoric acid derivatives (Fig. 3) [194]. Compound 350 shows high selectivity towards adenosine triphosphate (ATP) over other tested phosphoric acid derivatives: monophosphate, pyrophosphate, triphosphate, adenosine monophosphate and adenosine diphosphate. Moreover, the absorption spectrum of 350 responds to ATP specifically even if other analogues existed. The Cu2+ complex of compound 350 forms 1:1 complex towards ATP with binding constant 6640 M−1, accompanied with remarkable blue shift in the absorption spectrum. 1H NMR experiments revealed that adenine moiety of ATP existed in the γ-CD cavity. Dipicolylamine moiety provides additional recognition site for phosphoric groups of ATP. The cooperation of these interactions probably results in the high selectivity to ATP.

Fig. 3
figure 3

Reproduced with permission from Ref. [194]. Copyright 2017 from American Chemical Society

a Selective sensing to ATP by Cu·350 in the presence of phosphoric acid derivatives. b Suggested structure of Cu·350/ATP complex.

In 2006, Chun et al. reported a nitroazophenolic crown ether 219 (see Scheme 6) with asymmetric centers for enantiomeric recognition of amines [135]. Spectrophotometric method verified that the binding of amines resulted in changes of the absorption spectrum of the host, and that two enantiomers formed complexes with different stabilities. The recognition process can also be monitored by cyclic voltammetry. The oxidation of the phenol group of 219 shows a single peak at voltammograms. The addition of an alkyl amine is accompanied by a new oxidation wave at potentials that are different for each enantiomer. The difference between the oxidation potentials of the complexes formed by 219 with two enantiomers of 2-amino-2-phenylethanol is 44 mV. This method allows quantifying the ratio of enantiomers in an R/S mixture because the peak potential of 219 varied linearly with the enantiomer ratio.

In 1998, Kim and Chang reported a calix[4]arene chromophore 928 (see Scheme 18) with two distal azophenol moieties [391]. NMR data showed 928 adopted a syn-oriented conformation. Binding of Ca2+ by 928 was determined by extraction between CaCl2 solution at pH 7 and CHCl3 contain host 928. A bathochromic shift from 437 to 605 nm was observed (Fig. 4) and the solution color changed from yellow to greenish blue. Other similar cations induced no shift at comparable concentrations and the selectivity to Ca2+ over them were 195 (Sr2+), 725 (Mg2+) and 680 (Na+). The nitrogen atoms on 928 are important for both binding at neutral condition and Ca2+ selectivity.

Fig. 4
figure 4

Reproduced with permission from [391]. Copyright 1998 from American Chemical Society

UV absorption spectra of 928 upon extraction with aqueous metal chloride solution.

In 2012, Pulpoka and co-workers described a selective fluoride sensor based on an azo calix[4]arene-strapped calix[4]pyrrole 473 (see Scheme 17) [269]. The color modulation of CH3CN solution of 473 was observed in the presence of certain anions, which was also reflected in distinctive absorption changes. Upon addition of six equivalents of F, a bathochromic shift from 395 nm to about 600 nm in the absorption spectrum of 473 was observed. Other tested anions only imparted much smaller, if not negligible, absorption changes. The spectroscopic changes upon basic anions complexation were ascribed to the modulation of charge transfer from the oxygen of azophenol moieties to the nitrophenyl chromophores. Spectroscopic analyses showed the formation of both 1:1 and 1:2 complexes for F. By adding Ca(NO3)2 into a solution of the 473·F complex, the F ions were salted out in the form of CaF2, which resulted in the recovery of the original orange color of free receptor 473 (Fig. 5). In contrast, when Ca(NO3)2 was added to the corresponding CH3COO, BzO, or H2PO4 complexes, ion-pair complexes were formed without noticeable color changes.

Fig. 5
figure 5

Reproduced with permission from Ref. [269]. Copyright 2011 from American Chemical Society

F coordination by 473 and displacement by Ca(NO3)2.

In 1985, Kaneda et al. developed a series of novel spherand azophenol dyes 1236 (see Scheme 26), displaying Li+-specific colouration [560]. The crystal structure of 1236a (Fig. 6), an analogue of 1236, showed that the radius of the cavity is 0.68 Å. This radius is compatible with the ionic radius of Li+ (0.60 or 0.73 Å) and is smaller than that of Na+ (0.95 or 1.16 Å). In chloroform, only lithium salt in 63 kinds of salts leads to the change in absorption spectrum in the presence of excess piperidine as base, with a remarkable color change from yellow to violet. The “perfect” selectivity is obtained due to the steric effect caused by the narrow entrance to the cavity rejects larger cations.

Fig. 6
figure 6

Reproduced with permission from Ref. [560]. Copyright 1985 from American Chemical Society

Molecular structure of 1236a shown by the ORTEP drawing.

In 1988, the same group studied the binding properties of azophenol crown ether 255 (see Scheme 6) containing a benzoic acid moiety [116]. The solution of 255 was yellow and showed absorption maxima at 400 nm in chloroform. Large red shift of its UV–Vis spectrum could be observed when 255 formed complexes with amines. The solution color turned to blue with monoamines while to pink with diamines, indicating a very different coloration based on the two ionic binding sites in the crown cavity. Titration experiments show that 255 forms a 1:1 complexes with some equivalents of diamines and 1:2 complexes with large excess of diamines. The molecular structure of the salt complex between 255 and piperazine was determined by X-ray crystal analysis (Fig. 7). In the 255·piperazine complex, the protonated piperazine is sandwiched between phenolate and benzoate planes in “chair” form. Strong hydrogen bonds are formed between the phenolic oxygen atoms of 255 and the piperazine guest.

Fig. 7
figure 7

Reproduced with permission from Ref. [116]. Copyright 1988 from American Chemical Society

An ORTEP drawing of the molecular structure of the 255·piperazine complex.

In 2011, Rebek and co-workers synthesized photoswitchable cavitands 977 and 978 (see Scheme 19) bearing the azo moiety as one of the walls [415]. The thermal- and photo-isomerization behaviors of azo cavitands 977 and 978 were analyzed in d12-mesitylene (Fig. 8). Transcis photo-isomerization is achieved by illuminating with 365 nm light for 15 min to reach a photostationary state. Cistrans conversion can be achieved by heating to 164 °C for 5 min or irradiating with light longer than 450 nm for 20 min. Transcis and cistrans cycles can be repeated 5 times without degradation of the system. Both the trans-isomers of 977 and 978 are able to bind neutral adamantane guests. The binding behaviors of trans azo cavitands 977 and 978 were studied in d12-mesitylene. The highest values of stability constants were observed for 1-adamantanecarbonitrile and 2-adamantanone, which can interact with the amide on the upper rim of cavitands. The complexation of adamantane guests can be light-controlled for 978, but not for 977 without tert-butyl substitution. This striking contrast is because the bound guest is replaced with an introverted tert-butyl.

Fig. 8
figure 8

Reproduced with permission from Ref. [415]. Copyright 2011 from Royal Society of Chemistry. (Color figure online)

1H NMR spectra in d12-mesitylene showing trans-978 (bottom) and self-contained cis-978 (top). The inverted tert-butyl signal is marked by a blue circle.

In 2016, Scherman and co-workers described the synthesis and photo-responsive binding properties of macrocycle 1220 containing bis(imidazolium)-azobenzene motifs (Fig. 9) [553]. NMR analysis revealed macrocycle 1220 existed as all-trans stereoisomer in a freshly prepared trideuterioacetonitrile solution. Isomerization products generated by irradiating UV light were a mixture of stereoisomers E,E-1220 (18%), E,Z-1220 (38%) and Z,Z-1220 (44%). Visible light illumination causes the photostationary state E,E-1220 as the main isomer. Stereoisomer E,E-1220 could complex with 4,4′-dipyridyl-N,N′-dioxide (4DPDO), 4-phenylpyridine N-oxide and biphenyl-4,4′-dicarboxylate. Irradiation of the complex with UV light induced the trans- to cis-isomerization of the azo groups, resulting in the 4DPDO guest release from the cavity. Further exposure of the mixture to visible light reverts the guest encapsulation into the macrocycle cavity.

Fig. 9
figure 9

Reproduced with permission from Ref. [553]. Copyright 2016 from Wiley-VCH Verlag GmbH & Co. KGaA

Schematic representation of the photo-controlled catch and release of 4DPDO by 1220.

Self-assembly

In 2008, Liu et al. used click chemistry to covalently connect the host and guest in threaded β-CD-based azobenzene complexes (Fig. 10) [172]. Azobenzene substituted with propargyl alcohol was coupled to azide functionalized β-CD, forming different isomers depending on the reaction conditions. The hydrothermal 1,3-cycloaddition led to self-locked isomer 334a (see Scheme 10), whereas Cu(I) catalyzed click chemistry gave the self-unlocked isomer 334b. A possible mechanism for the synthesis of 334a involved the primary formation of the inclusion complex of β-CD with azobenzene, and the subsequent reaction between the azido group of β-CD with the ethynyl group of azobenzene. In aqueous solution and in solid state, 334a formed dimer capsule, in which the azobenzene moieties were included by both its own cavity and the cavity of the second β-CD, as demonstrated by ROESY-NMR and solid state structure. In DMSO, however, 1H NMR spectroscopy shows that 334a existed as the monomeric self-locked conformer, which could be regarded as a new type of [1] rotaxane without a stopper. In water–DMSO mixture, 334b assembled to linear oligomeric supramolecule, as demonstrated by high resolution TEM and FTICR-MS. The above study shows that the spatial arrangement factor can be of great importance for the self-assembling morphology.

Fig. 10
figure 10

Reproduced with permission from Ref. [172]. Copyright 2008 from American Chemical Society

Deduced self-assembly structures of 334a (top) and 334b (bottom).

In 2015, the same group reported the photo-controlled nanotube–nanoparticle conversion for the assembly of an amphiphilic porphyrin derivative and azobenzene-bridged bis(permethyl-β-CDs) 365 (see Scheme 11 and Fig. 11) [207]. The porphyrin derivative could form aggregates with an average hydrodynamic diameter about 180 nm. Upon addition of trans-365 in solution of the porphyrin derivative, the morphology of the aggregate transformed to secondary assembly nanotubes, which is revealed by TEM images. The average inner and outer diameters of the nanotubes were about 45 and 61 nm with a wall thickness of about 8 nm. Upon photoisomerization the nanotubes reversibly changed to nanospheres with the average diameter of 180–220 nm.

Fig. 11
figure 11

Reproduced with permission from Ref. [207]. Copyright 2015 from Wiley-VCH Verlag GmbH & Co. KGaA

Illustration of the secondary assembly of amphiphilic porphyrin derivative mediated by trans-365.

In 2013, Aseyev and co-workers synthesized polymer 443 (see Scheme 15) with azo calix[4]arene in the main chain that was substituted with tetraethylene glycol monomethyl ether chains [254]. The poly(azo calix[4]arenes) exhibited lower critical solution temperature behavior in aqueous solution and upper critical solution temperature behavior in alcohols. In alcohols, the phase transition temperature of the polymers decreases proportionally with increasing the cis content by irradiating with 365 nm light. The cloud point can be modulated by irradiation or thermal relaxation depending on the trans/cis ratio of the azobenzene groups. The 365 nm light assisted writing on solutions of the polymer in alcohols can be realized (Fig. 12).

Fig. 12
figure 12

Reproduced with permission from Ref. [254]. Copyright 2013 from American Chemical Society

Photographs of poly(azo calix[4]arenes) 443 in ethanol: a at 20 °C before irradiation; b after photoassisted writing at 20 °C; c at 40 °C and d at 20 °C after thermal relaxation.

In 2016, Ballester and co-workers developed a heterodimeric capsule containing an azobenzene-appended calix[4]arene 638 (see Scheme 18) and a tetraurea calix[4]pyrrole without azobenzene 638a in the presence of polar guests [289]. In the presence of trimethylphosphine oxide guest, irradiating the heterodimeric capsule results in the formation of all-cis-638·638a capsule, indicating photo-isomerization in only half of capsule is not sufficient to cause the disassembly of the capsule. In a more complicated system, in which both trimethylphosphine oxide and bis-(N-oxide N,N,N′,N′-tetramethylamino)hexane existed, two capsules, 638·638a and 638a·638a, are formed at the same time and the equilibrium between them can be modulated by light.

In 2014, the same group further developed a tetraurea aryl extended calix[4]pyrrole 1007 (see Scheme 20) with four azobenzene groups at the upper rim [437]. By encapsulating 4,4′-bipyridine bis-N,N′-oxide as a template, 1007 assembles into a hydrogen-bonded dimeric capsule (Fig. 13). The capsule can be detected by 1H NMR spectroscopy in solution only when all azobenzene units are in trans forms. The capsule disintegrates when irradiated with 400 nm light and reassembles after cis-to-trans relaxation in the dark. They further presented photo-responsive homodimeric and heterodimeric capsules using two novel calix[4]pyrrole [433]. This study may be further developed to photo-responsive drug delivery if can operate in aqueous media.

Fig. 13
figure 13

Reproduced with permission from [437]. Copyright 2014 from Royal Society of Chemistry

Structures of arylextended tetraurea calix[4]pyrrole and bis-pyridine-N-oxide (left) and energy minimized structure of the capsular assembly (right).

In 2008, Gin and co-workers prepared a shape-switchable metallacycle 1259 (see Scheme 27) bearing azobenzene units via coordination-driven self-assembly [572]. The Pt containing metallacycle forms mesogen at room temperature in a thermotropic columnar liquid crystal (LC) phase, which was swollen in polar solvents (dioxane, diglyme, and ethylene glycol) to further form a lyotropic LC phase (Fig. 14). For the lyotropic LC state of unpolymerised 1259, irradiation with 375 nm light resulted in conversion to the cis-isomer, accomplished by phase transition from a liquid crystalline phase to an isotropic phase. However, cross-linking the lyotropic LC assembly led to a nanostructured polymer network, preventing disruption of the LC order when photoconversion to the cis-isomers upon UV irradiation. The material may be useful to drug-delivery or controlled nanofiltration applications.

Fig. 14
figure 14

Reproduced with permission from Ref. [572]. Copyright 2008 from Wiley-VCH Verlag GmbH & Co. KGaA

Proposed, idealized model of the thermotropic columnar hexagonal LC phase of 1259 (left). Addition of polar solvent serves to swell this phase (right).

In 2011, Norikane et al. developed photo-responsive liquid crystallines based on azobenzenophanes 1021 and 1022 (see Scheme 21 and Fig. 15) [481]. The self-assembling material can be switched by light. In the liquid crystalline state, dimer 1021 behaves as a rod-like molecule exhibiting smectic phases, while the disk-like trimer 1022 forms columnar phases. These materials exhibit isothermal phase transition from liquid crystalline to isotropic state upon UV light irradiation because of shape change of the molecules. The liquid crystalline phase forms again upon heating the isotropic phase at 120 °C for 10 s. Such photo-responsive materials could be useful in applications to photolithography and photo-responsive adhesives.

Fig. 15
figure 15

Reproduced with permission from Ref. [481]. Copyright 2011 from Royal Society of Chemistry

Polarizing optical photomicrographs of 1021 at 120 °C (ac) and 1022 at 25 °C (df) before 365 nm light UV irradiation (a, d), during irradiation for 5 s (b, e) and 10 s after ceasing the irradiation (c, f). The initial states are SmC (a) and columnar (d) phases. The dark area in (bf) corresponds to an isotropic phase.

In 2009, Wang and co-workers integrated four azobenzene units in the backbone of azobenzenophane 1015 (see Scheme 21) by connecting through intermediate carbazole groups [474]. Although 1015 does not form an extended 2D network on its own, self-assembled monolayers of host–guest networks were observed by mixing it with 1,3,5-tris(10-carboxydecyloxy)-benzene (TCDB) on a highly ordered pyrolytic graphite surface. Depending on the mixing ratio of TCDB and the macrocycle, the TCDB network captures either a monomer or a dimer of 1015. The structures of the networks dependent on the isomers of 1015 can be distinguished by scanning tunneling microscopy, which revealed that all the azo groups are in the trans form (t,t,t,t). Upon irradiation of the monolayer by 366 nm light, azobenzene units of the macrocycle partially isomerize and give rise to different isomers including transtranstranscis (t,t,t,c) and transcistranscis (t,c,t,c) isomers (Fig. 16).

Fig. 16
figure 16

Reproduced with permission from Ref. [474]. Copyright 2009 from American Chemical Society

Molecular models of the conformational and photoisomers of 1015 in the photoisomerization.

In 2013, Wegner and Reuter reported an azobenzene macrocycle 1223 (see Scheme 25) displaying the light-controlled switchable π-stacking (Fig. 17) [482]. 1223 can form 3D networks through significant π-stacking interactions. Light-controlled sol–gel transition of 1223 could be observed only in aromatic solvents, because they are incorporated inside the 3D π-stacking network. These results suggested that 1223 might be useful for building up materials for small molecules controlled release by light stimulation.

Fig. 17
figure 17

Reproduced with permission from Ref. [482]. Copyright 2013 from Royal Society of Chemistry

Schematic representation of 3D-networks dimeric switchable azobenzene macrocycle 1223. In the all-E conformation linked p-stacks are formed (left). Irradiation induces isomerization to a mixture of Z-isomers all deviating from planarity, which impedes the supramolecular interaction (right).

In 2017, Xie et al. synthesized a carbohydrate-based macrocyclic azobenzene 1272 (see Scheme 27) using one-pot O-alkylation mediated macrocyclization approach [574]. 1272 exhibits reversible photo-isomerization between trans-and cis-isomers upon irradiation of light with different wavelengths. Moreover, thermal stability of the cis-isomer (t1/2 = 51 days) is better than of the acyclic analogue (t1/2 = 19 days). Circular dichroism study revealed the chirality transfer from the sugar unit to azobenzene, which adopts preferentially P-helicity for both trans-and cis-isomers. In cyclohexane and ethanol, 1272 can form organogels, which are responsive to temperature, light and mechanical force (Fig. 18). Furthermore, 1272 displays a helical transition behavior during gelation that can be modulated by changing temperature.

Fig. 18
figure 18

Reproduced with permission from Ref. [574]. Copyright 2017 from Wiley-VCH Verlag GmbH & Co. KGaA

Pictures showing the multi stimuli-responsive behavior of gel of 1272 in cyclohexane.

Application

Based on the exploration of recognition, self-assembly and other properties, the chemistry of azo macrocycles gives rise to several applications. Here, we highlight a few of them containing selective sensing and imaging, photocontrolled supramolecular catalysis, molecular machine, and so on.

Azo isomerization mechanism

In 1982, Rau and Lueddecke synthesized two azobenzenophanes 1049 and 1069 (see Scheme 21) for application in studying transcis-isomerization mechanism of azobenzenes [484]. For the photoisomerization, two different pathways have been proposed: a rotation around the N–N double bond and a direct inversion of the orientation of one of the phenyl rings. For the two macrocycles, inversion was the only possible isomerization pathway because of the steric hindrance. The quantum yield of the transcis-isomerization was 0.24 for n–π* excitation of 1049, which is similar to that of azobenzene (0.23). This indicates that the transcis-isomerization of azobenzene under n–π* excitation is through an inversion mechanism.

Sensing and imaging

Although some works involved in sensing have been described in the section of recognition, we herein discuss sensing for works closer to practical application. In 1992, Kumar and co-workes reported a small-cavity cryptand phenol 319 (see Scheme 9) which could be applied for the colorimetric determination of lithium in blood serum without sample pretreatment or solvent extraction [161]. The compound 319 exhibits a greater than 4000:1 selectivity for lithium over sodium due to rigid configuration of a well-preorganized binding site for lithim complexation. The standard curve for lithium is linear up to 3.5 mM and contains the therapeutic range. Serum was diluted 40 times for the measurement and the obtained results for serum spiked with lithium correlated well with flame photometry measurements.

In 2010, Lang and co-workers developed a pH sensor 549 (see Scheme 18) based on azo calix[4]arene bearing four carboxyl groups [319]. The crystal structure obtained in DMF solution revealed that 549 adopt a flattened cone conformation. Moreover, the aqueous solution of 549 displayed distinct color change responding to pH in the range of 1–13, which is broad than most of pH sensing compounds (Fig. 19). The color change is reversible, suggested that 549 may have a potential application as a good pH sensor.

Fig. 19
figure 19

Reproduced with permission from Ref. [319]. Copyright 2010 from Wiley-VCH Verlag GmbH & Co. KGaA. (Color figure online)

a UV–Vis spectra of 549 titrated with a dilute NaOH solution. b Color changes corresponding to different pH values.

Peoples are interested in concentrations of biological molecules, not only in tissue and body fluid but also in single cell level, in which fluorescence imaging is highly on demand. In 2016, Yilmaz and co-workers developed anthracene and pyrene appended azo calix[4]arene 666 and 667 (see Scheme 18) and applied 666 in fluorescence imaging of Cu2+ in living cells [355]. The azo calix[4]arenes shown selective complexation to Cu2+ over other metal ions, which is revealed by UV–Vis and fluorescence spectroscopies. The binding constants are calculated as 2.34 × 103 and 7.51 × 103 M−1 for 666 and 667, respectively. Once incubated with 666 containing medium following Cu2+ treatment, the tested SW-620 cells showed obvious fluorescence (Fig. 20).

Fig. 20
figure 20

Reproduced with permission from Ref. [355]. Copyright 2016 from Elsevier

Fluorescence image and their corresponding bright-field transmission images. a Control cells (only cells without Cu2+or ligand 666); b fluorescence image of SW-620 cells after treatment with Cu2+and ligand 666; c merge view of bright field and fluorescence images.

Devices

In 1992, Tamaoki et al. synthesized and studied the isomerizations of azobenzenophane 1013 (see Scheme 21) in polycarbonate films through spectrophotometry [462]. The efficiency of photoisomerization of 1013 is dependent on the light intensity both in solution and on polymer film. This unique feature can be applied in the construction of photochromic memory with non-destructive read-out property (Fig. 21). The memory was written and read-out by exposure to 366 nm light at 5 mW cm−2 and 20 µW cm−2, respectively. The memory can be initialized by irradiating with light of wavelength longer than 450 nm.

Fig. 21
figure 21

Reprinted from Ref. [462]. Copyright 1992 with permission from Elsevier

Procedures for writing, monitoring and erasing of memories.

In 2006, Rouis et al. elaborated organic diodes based on sandwiching azo calix[4]arene layers between indium-thin oxide and Al contacts [282]. Optical and electrical properties of these devices were studied by current–voltage characteristics and electrical impedance spectroscopy in a wide frequency range. These materials displayed characteristic of semiconductors with energy band gaps in the range of 1.7–3 eV and the electrical characteristics follow a space charge limited conduction behavior. Authors also determined and fitted the relaxation frequencies of the hopping charge carriers.

Photocontrolled supramolecular catalysis

In 2003, Cacciapaglia et al. devised a photoswitchable supramolecular catalyst based on Ba2+ complex of azo bis(benzo-18-crown-6) ether 176 (see Scheme 5) to photocontrol the rate for ethanolysis of 4-carboxyacetanilides in basic solution (Fig. 22) [109]. Efficient ethanolysis requires the cooperation of two metal centers: one to bind the carboxylate group, and the other to deliver the alkoxide to the amide. The trans-azobenzene spacer is too extended to allow for proper two-point binding and cis-176 is efficient catalyst (kcis/ktrans up to 5), which is confirmed by molecular models. The catalytic activity could be reversibility switched in situ between ‘faster’ and ‘slower’ for several times by alternating exposure to UV or visible light.

Fig. 22
figure 22

Reproduced with permission from Ref. [109]. Copyright 2003 from American Chemical Society

Computer-generated structures of interswitchable trans and cis forms of 176·2[Ba2+].

Molecular machine

In 2006, Aida and co-workers achieved the coupling of several molecular motions in a light-powered system (Fig. 23) [507]. A photoresponsive host contain an azobenzene unit, a ferrocene unit as pivot, and two porphyrin units as binding sites. The host complexes with 4,4′-biisoquinoline guest with high binding constant. In the formed complex, the change in molecular shape on photoisomerization of the azobenzene is transmitted via rotation of ferrocene and ultimately induced the rotary motion in the guest. The author confirmed the motion by studying the chirality of the guest through circular dichroism spectra. The original state can be regenerated through irradiation with visible light. The system convert light energy into significant motion and larger system that could address remote control of molecular motion might be developed through similar strategy.

Fig. 23
figure 23

Reproduced with permission from Ref. [507]. Copyright 2006 from Nature Publishing Group

Design and concept of light-powered molecular pedal.

Others

In 2010, Haberhauer and Kallweit described the unidirectional photoisomerization process of in an azobenzene bearing macrocycle 1267 (Fig. 24) [575]. Trans- to cis-isomerization of azobenzene generates two enantiomeric cis-isomers with P or M chirality, respectively. For most of azobenzene derivatives the two enantiomeric isomers are energy equal and are in fast racemization. It is possible to switch the azobenzene unit unidirectionally by linking a chiral clamp. The irradiation with light of the achiral trans-isomer of 1267 gives rise to the cis-isomer with P helicity. The quantum mechanical computation revealed that the cis-(P) isomer of 1267 is much more stable than the cis-(M) isomer.

Fig. 24
figure 24

Reproduced with permission from Ref. [575]. Copyright 2010 from Wiley-VCH Verlag GmbH & Co. KGaA

Unidirectional switching of the cyclic azo macrocycle 1267.

In 2013, Deligöz and co-workers synthesized four azo calix[4]arene dyes for the application in dyeing fibers, including cotton, wool, acetate, polyether and polyamine [322]. Both perspiration fastness and water fastness of fibers dyed by these compounds are good. Furthermore, thermogravimetric analysis revealed that these dyes are all stable up to 245 °C, which can be effectively used in most kind of inks. It is worth mentioning that the tautomerization of these macrocycles need to be considered in the future.

Summary

Introducing azo groups into macrocycles can endow them with more intriguing chemical and physical properties, such as efficient recognition, guest induced color change for optical sensing, and photo-response. In this review, we first comprehensively summarized the structures of various macrocycles containing azo group. Most of popular macrocycles have many azo-modified derivatives, especially for crown ether and calixarene, due to the convenient synthesis of azo compounds and the facile modification of these macrocyclic scaffolds.

Furthermore, we reviewed several representative examples of these azo macrocycles for molecular recognition, self-assembly and application to give readers an impression of their properties and research status. In general, most works about azo macrocycles focused on their syntheses and (photo-responsive) model guest recognition. We listed > 1200 azo macrocyclic molecules but rarely found the works about application, and even self-assembly.

In our opinion, the supramolecular chemistry and application of azo macrocycles are still in their infancy and will develop flourishingly in the future since both azo compounds and macrocycles have high status in various fields. The breakthrough relies on finding methods to use both properties of azo compounds, such as photo- and redox-response, fast internal conversion and color, and unique supramolecular chemistry of each macrocycle. We have a vision that azo macrocycles will have more fascinating application in fields of diagnosis and therapy, molecular machine, liquid crystal and nonlinear optics. Its biomedical application, for example, tumor-selective imaging and targeting drug delivery based on hypoxia-responsive azo macrocycle, is of our particular interest in the near future.