Next Article in Journal
Two Novel Lyso-Ornithine Lipids Isolated from an Arctic Marine Lacinutrix sp. Bacterium
Next Article in Special Issue
Azocalix[4]arene-Rhodamine Supramolecular Hypoxia-Sensitive Systems: A Search for the Best Calixarene Hosts and Rhodamine Guests
Previous Article in Journal
Automatic Identification of Analogue Series from Large Compound Data Sets: Methods and Applications
Previous Article in Special Issue
Molecular Recognition of Imidazole Derivatives by Co(III)-Porphyrins in Phosphate Buffer (pH = 7.4) and Cetylpyridinium Chloride Containing Solutions
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Macrocyclic Receptors for Identification and Selective Binding of Substrates of Different Nature

by
Galina Mamardashvili
,
Nugzar Mamardashvili
* and
Oscar Koifman
G.A. Krestov Institute of Solution Chemistry of the Russian Academy of Sciences, Akademicheskayast. 1, 153045 Ivanovo, Russia
*
Author to whom correspondence should be addressed.
Molecules 2021, 26(17), 5292; https://doi.org/10.3390/molecules26175292
Submission received: 15 May 2021 / Revised: 4 August 2021 / Accepted: 18 August 2021 / Published: 31 August 2021
(This article belongs to the Special Issue Molecular Recognition of Host/Guest Molecules)

Abstract

:
Molecular recognition of host/guest molecules represents the basis of many biological processes and phenomena. Enzymatic catalysis and inhibition, immunological response, reproduction of genetic information, biological regulatory functions, the effects of drugs, and ion transfer—all these processes include the stage of structure recognition during complexation. The goal of this review is to solicit and publish the latest advances in the design and sensing and binding abilities of porphyrin-based heterotopic receptors with well-defined geometries, the recognition ability of which is realized due to ionic, H-bridge, charge transfer, hydrophobic, and hydrophilic interactions. The dissection of the considered low-energy processes at the molecular scale expands our capabilities in the development of effective systems for controlled recognition, selective delivery, and prolonged release of substrates of different natures (including drugs) to their sites of functioning.

1. Introduction

Molecular recognition is a process in which host molecules (receptors) select and bind guest molecules (substrates) into structurally highly organized systems through multipoint intermolecular interactions [1,2,3,4,5,6,7]. Porphyrins are extremely useful for this purpose. Their unique properties are the result of the unusual geometric and electronic structure of the porphyrin macrocycle, containing a developed aromatic conjugated system. Selective chemical modification of porphyrins by fragments of other classes of compounds makes it possible to synthesize molecular systems that are different in their nature and positions of the reaction centers relative to each other [8,9,10,11,12,13,14]. This review discusses the progress made so far in the design and sensing and binding abilities of porphyrins modified with calix[4]arenes, calix[4]pyrroles, oxacalix-[2]-arenes-[2], and resorcinarenes and with bulky carbazolylphenyl substituents of different generations towards substrates of different natures. The molecular structures, sensing mechanism, and successful applications of these macrocyclic compounds with well-defined geometries as heterotopic receptors for molecular recognition and reversible binding of anions, cations, and small organic molecules are discussed in detail. Data on the ion-assisted binding ability of tetrapyrrole macroheterocyclic compounds decorated with additional complexing cavities and channels are of particular interest. It is shown that the binding ability of the porphyrin core depends on the conformation mobility of the substituents and the possibility of formation of intramolecular cavities and channels of various shapes within them. We hope this review will bring more attention to this rising field and inform researchers in related directions of its potential applications.

2. Molecular Receptors Based on Porphyrin Conjugates with Macrocyclic Compounds of Different Natures

Interest in the chemistry of porphyrins was first sparked by the participation of these compounds in photosynthesis. However, it has now turned in another direction as porphyrins can be used as selective molecular receptors for certain types of substrates [15,16,17,18,19,20,21,22,23,24,25,26,27,28,29,30,31,32,33,34,35]. The fact that it is possible to chemically modify the tetrapyrrole molecule has enabled researchers to synthesize porphyrin conjugates with other macrocyclic compounds(porphyrin conjugates with calix[n]arenes [36,37,38,39,40,41,42,43,44,45,46,47,48,49,50,51,52,53,54], cyclodextrines [55,56,57,58,59,60,61,62,63,64,65,66], calix[n]pyrroles, fullerenes, ferrocenes, and resorcinarenes [66,67,68,69,70,71,72,73,74,75,76,77,78,79,80]) possessing their own properties, allowing them to form complexes with ions and organic molecules of different natures. Table 1 and Table 2 present some data on “molecular tweezer” macrocyclic receptors and the substrates they bind, based on results of the analysis of the most interesting works in this field.
Especially interesting among such compounds are covalently bound porphyrin-calix[4]arene macrocyclic receptors [8,10,36,37,38,39,40,41,42,43,44,45,46,47,48,49,50,51,52,53,54]. The first calix[4]arene-porphyrin conjugate capable of selectively identifying small organic molecules was described in [47].
By making ortho-carbomethoxyaldehyde (1) interact with pyrrole in the presence of trifluoroacetic acid, α-unsubstituted meso-aryl dipyrromethane (2) was synthesized (Scheme 1) which was thentransformedinto the respective 5,15-di-(carbomethoxyphenyl)-porphyrin (3) through condensation with benzaldehyde and subsequent oxidation of the intermediate porphyrinogen by 2,3-dichloro-5,6-dicyano-benzoquinone (DDQ). Hydrolysis of ester groups 3 subjected to the action of lithium hydroxide in tetrahydrofuran and subsequent interaction with oxalyl chloride in dichloromethane yielded porphyrin diacylchloride (4), which was later made to interact with the diamine of the respective calix[4]arene (5) in dichloromethane in the presence of catalytic amounts of triethylamine to produce the target porphyrin-calix[4]arene (6, M=H2). By boiling ligand 6 with zinc acetate in a methanol–chloroform mixture, the zinc complex of this compound (7, M=Zn) was obtained.
The 1H-NMR spectra of porphyrinate 7 in the presence of pyridine (Py) and imidazole (Im) have two rows of signals of aromatic protons belonging to the free and bound porphyrinate molecules of the substrate (Figure 1). The ring current shielding effect of the porphyrin macrocycle results in an up-field shift of the aromatic protons of the bound substrate molecules compared with the respective protons of the unbound substrate molecules.
The structure of the 7-Py complex determined by the molecular mechanics methods using the MMFF force field of the MacroModel (V6.5) software package on the SGI O2 workstation is shown in Figure 2. Applying the method of spectrophotometric titration, it was established that porphyrinate 7 is most capable of forming complexes with imidazole (Table 3).
Thus, Zn-porphyrin 7 capped with calix[4]arene offers in its cavity a guest-binding environment for selective binding of small organic molecules, which makes it possible to distinguish even small structural differences between pyridine and 4-methylpyridine.Guests that are shaped in a way that is complementary to the confined cavity are capable of fitting into and binding to the cavity by van der Waals attractive interactions.
Another example of a calix[4]arene-capped porphyrin (8) was obtained by the authors of [48] on the basis of tetra-tert-butyl-tetrakis-(hydroxycarbonylmethoxy)-calix[4]arene (9) and tetra-(2-aminophenyl)-porphyrin (10) according to Scheme 2.
A zinc complex of compound (11) was obtained by treatment of ligand 8 with zinc acetate in DMF at room temperature. It was established that the modified porphyrin (3) shows very interesting ion-binding ability (Scheme 3). In NaClO4, Na+ is bound to the calix[4]arene cavity and ClO4 exists as a counter anion to give the 429 nm band, whereas in KI, K+ is bound to the calix[4]arene cavity and I coordinates to Zn(II) to give 446 nm. These results indicate that the cavity with a hard-soft ditopic binding site is very suitable for recognition of I ions.
The synthesis and binding ability of a Zn-porphyrin (12) with a calix[5]arene cap were presented in [49]. In order to synthesize such a bridged porphyrin, 5,15-Bis-(carboxynaphtyl)-10,20-diphenylporphyrin (13) with syn arrangement of two naphthalene rings was prepared. Reaction of the porphyrin 13 with calix[5]arene with two amino functions (14) and subsequent treatment of the reaction product with zinc acetate gave the calix[5]arene-capped Zn-porphyrin 12 (Scheme 4).
It was shown that the calix[5]arene-capped Zn-porphyrin 12 is able to bind 4-methylpyridine much stronger that non-modified Zn-tetrphenylporphyrin due to the accommodation of the guest molecule into the calix[5]arene cavity. On the other hand, bulky substituents on the pyridine group, as in the case of 4-tert-butylpyridine, 4-phenylpyridine, and 3,5-dimethylpyridine, prevent the guest molecules’ penetration into the calix[5]arene cavity, and the constants of their binding to the modified receptor are commensurate with the corresponding binding constant to the Zn-tetraphenylporphyrin.
For the purpose of developing a novel photoelectronic active porphyrin derivative with good solubility and suitable selective recognition ability for the guest neurotransmitter dopamine (DA) molecule, a functionalized porphyrin 15 was designed (Scheme 5) by the reaction between 5-iodinephenyl-10,15,20-tris-(1-pyrene)-porphyrin 16 and propinyl-calix[4]arene in mixed trimethylamine/dichloromethane(1:1) solvent in the presence of PdCl2(PPh3)2 [10].
It was shown that a combination of the effective molecular recognition/enrichment and good optoelectronic properties derived from calix[4]arene and pyrene substituents as well as the p-conjugated porphyrin skeleton leads to the fluorometric and voltametric dual-mode sensing activity of the calix[4]arene-porphyrin 8 for DA in the range of 1–50 mM and 2–100 mM, respectively, with remarkable specificity, representing the first multi-modal molecular sensor for DA.
“Molecular tweezers” based on the thiacalix[4]arene-porphyrin conjugate 17 capable of selectively identifying the C70fullerene were obtained by the authors of [36].
The procedure of synthesis of the tetra-formyl derivative of thiacalix[4]arene in the 1,3-alternate conformation is presented in Scheme 6. Bromation of commercially available tetrahydroxy-thiacalix[4]arene (18) by N-bromosuccinimide in acetone was used to synthesize tetrabromo-derivative (19), which was then turned into the respective propoxy-derivative (20) through its interaction with propyl iodide in the presence of potassium carbonate in acetone. It should be said that introduction of bulky propoxy-substituents into the upper ring of the initial thiacalix[4]arene makes the molecule change its cone conformation to the 1,3-alternate one. The interaction of (20) with tert-butyllithium in tetrahydrofuran accompanied by cooling allowed preparation of the target thiacalix[4]arene tetraaldehyde (21), which was then condensed with monoamino-porphyrin (22) in dichloromethane in the presence of molecular sieves, to synthesize thiacalix[4]arene-Bis-(porphyrin) (17) (Scheme 7).
The ability of 17 to act as a receptor of the C60 and C70 fullerenes was studied by the authors using the 1H-NMR titration method. It was found that the C70 fullerene presence makes the signals of the porphyrin NH-protons shift up-field on the 1H-NMR spectrum. As the titration curve shows (Figure 3), the interaction leads to the formation of a 1:1 complex. It is shown (Table 4) that receptor 13 is highly selective to the C70 fullerene.
A calix[4]arene scaffolding was used by the authors of [38] to construct Bis-porphyrin (“jaws” porphyrin) hosts 2328 (Figure 4) for supramolecular binding of fullerene guests. Fullerene affinities were optimized by varying the nature of the covalent linkage of the porphyrins to the calix[4]arenes.
Binding constants for C60 and C70 in toluene were explored as a function of substituents at the periphery of the porphyrin, and 3,5-di-tert-butylphenyl groups gave rise to the highest fullerene affinities (26.000 M−1 for C60). The molecular modeling structure of 26-C60 is depicted in Figure 5. The origin of this high fullerene affinity has been traced to differential solvation effects rather than to electronic effects. Studies of binding constants as a function of solvent (toluene < benzonitrile < dichloromethane, cyclohexane) correlate inversely with fullerene solubility, indicating that desolvation of the fullerene is a major factor determining the magnitude of binding constants. A direct relationship between supramolecular binding of a fullerene guest to a Bis-porphyrin host and the appearance of a broad NIR absorption band has been established. The energy of this band moves in a predictable manner as a function of the electronic structure of the porphyrin, thereby establishing its origin in porphyrin-to-fullerene charge transfer.
The synthetic approach leading to the Pacman Bis-porphyrins was described by the authors of [51]. Established calixarene chemistry led to the diodideintermediate 29, which was isolated exclusively in cone conformation. Sonogashira coupling of 29 and the Ni-porphyrin ethynyl derivative (30) gave corresponding mono- (31) and Bis-porphyrin-calixarene (32) conjugates. It should be noted that both 3 and 4 display a cone conformation in solution as observed by 1H-NMR. Further alkylation of 32 took place at a rather slow rate to yield the conformationally frozen cone derivative 33 (Scheme 8).
The same authors developed a versatile stepwise synthetic approach allowing selection of the calix[4]arene conformation, the type of functionalization of the calixarene rims, and the anchoring point of the chromophores on the calix[4]arene spacer [39]. Connection of the ethynyl-derivative 36 with the calix[4]arene derivatives 34, 35, and 38 in the presence of 10% Pd(PPh3)2Cl2, CuI, and Net3 in toluene gave three calix[4]arene Bis-porphyrin conjugates 37, 39, and 42 (Scheme 9).
It was shown that receptors 37, 39, and 42 display very different behaviors in the presence of bidentate substrate DABCO, with 37 allowing a fast exchange between bound and free substrates, while 39 and 42 provide effective bounding of DABCO.
The works [40,41,53,54] are also devoted to the synthesis and analysis of the complexing properties of Bis-porphyrin-calixa[4]renes. Bis-porphyrin-calix[4]arene conjugate (43) (Figure 6) was synthesized similarly to Scheme 9 from Bis-iodo-calix[4]arene and nickel meso-ethynylporphyrinate by the Sonogashira method [53].
Works [40,41,54] describe the synthesis of calix[4]arene-Bis-(porphyrin) 44 (Figure 6) through condensation of calix[4]arene-Bis-(5,5′-unsubstituted dipyrromethane) (46) with 5,5′-diformyl-dipyrromethane in dichloromethane in the presence of catalytic amounts of trichloroacetic acid (Scheme 10), with subsequent oxidation of the intermediate porphyrinogen with dichlorodicyanobenzoquinone. By heating of porphyrin ligand (44) and zinc acetate in dimethylformamide, a zinc complex of receptor (45) (Figure 6) was obtained.
The synthesis of calix[4]arene-Bis-(porphyrin) 45 through condensation of diformylcalix[4]arene (48) with 4,4′-dimethyl-3,3′-diethyldipyrrolylmethane and benzaldehyde in dichloromethane in the presence of catalytic amounts of trichloroacetic acid is depicted in Scheme 11. By modifying the lower ring of calix[4]arene platform 45 and subjecting it to the action of pentaethylene glycol Bis-tosylate in the presence of cesium carbonate, we obtained a macroheterocycle (49) containing three functional components: porphyrin, calix[4]arene, and a crown ether fragment (Scheme 12a). By modifying the lower ring of the calix[4]arene platform 49 we obtained a macroheterocycle (50) containing two ethoxycarbonyl groups in the calix[4]arene platform (Scheme 12b).
By applying the 1H-NMR method, we found that formation of complexes of calix[4]arene-Bis-(porphyrin) receptors 45, 49, and 50 with bidentate triethylenediamine (DABCO) represents the 1:2 type of interaction. Scheme 13 illustrates the interaction of receptor 50 with DABCO as an example. This hypothesis is also confirmed by the separation of the signals of the nonequivalent DABCO protons in the 1H-NMR spectrum of the complex (Figure 7a). It means that the mismatch between the geometry of the interporphyrin complexing cavity of the receptor and the substrate leads to one-center binding of the studied bidentate ligand and formation of an external complex (51) with the stability constant Ka = 4.7 105 M−2 [49].
However, in the presence of a potassium cation, the formation of a complex of receptor 52 with DABCO represents a 1:1 interaction (Scheme 14). This hypothesis is also confirmed by the clear singlet of the equivalent protons of the substrate in the complex 1H-NMR spectrum (Figure 7b). It means that the conformational changes caused by the complex formation of the macrocycle crown ether fragment with the potassium cation make the porphyrin fragments of the Bis-porphyrin receptor get closer to each other and produce a two-center 1:1 complex between 52 and the organic substrate and a stable internal complex (53) with the stability constant Ka = 6.9 × 106 M−1.
Several different approaches for the synthesis of Bis-porphyrin-calixarenes have been developed [37,42]. First, Bis-porphyrin-calixarenes (5455), in which porphyrin moieties are linked to the calix[4]arene fragment through amide bridges, were synthesized by two methods: (i) by reaction of calix[4]arenes acyl chloride (56) with monoaminotetraphenylporphyrinininin the presence of base (Et3N); (ii) by a direct reaction of carboxylic acid (57) with monoaminotetraphenylporphyrin using dicyclohexylcarbodiimide as a coupling agent (Scheme 15) [42].
Later, Káset al. obtainedBis-porphyrin-calixarenescontaining directly linked porphyrin and calixarene fragments via condensation of an excess of pyrrole and p-methoxylbenzaldehyd with calix[4]arene-5,17-dialdehyde under BF3·Et2O catalysis in chloroform [37]. Subsequent oxidation of the intermediate porphyrinogen gave the corresponding Bis-porphyrin-calixarene conjugate (60) according to Scheme 16.
The 1H-NMR revealed the pronounced selectivity of the Bis-porphyrin derivative (60) towards C70 fullerene. Binding constants of the mono- (58) and Bis-porphyrin-calixarene (60) conjugates towards fullerene C70 are depicted in Table 5.
Calix[4]pyrrole was used in work [67] as the spacer holding the porphyrinate fragments in the face-to-face orientation to each other.
Thereaction of calix[4]pyrrole Bis-iododerivative (61) with zinc meso-ethynylporphyrinate (62) by the Sonogashira method was used to obtain calix[4]pyrrole-Bis-(porphyrin) conjugate (63) (Scheme 17). Applying spectrophotometric titration (Figure 8a), we established that formation of a complex of the calix[4]pyrrole fragment of receptor 61 with fluorine anions (Scheme 18) makes the Soret band in the absorption spectrum broader and less intensive, and in the region of the 1:1 molar ratio of the reagents, the reaction system becomes saturated (Figure 8b). It means that the interaction between receptor 63 and the anion produces 1:1 complexes (64) with the stability constant Ka = 15,000 M−1.
The down-field shift of the NH-protons of the calix[4]pyrrole fragment indicates that the complex formation causes the spacer to change its 1,3-alternate conformation to a cone (Figure 9). It is shown that complex formation involving a calix[4]pyrrole fragment can be used to control complex formation through the interporphyrin complexing cavity of receptor 63. Without the anion, the reaction of 63 with DABCO results in the formation of an external 1:2 complex (65) (Scheme 19) with the stability constant Ka = 4.7 × 105 M−2. The formation of external 1:2 complex 30 is confirmed by the splitting of the substrate proton signals in the 1H-NMR spectrum of the respective complex.
Some of the DABCO protons located closer to the macrocycle experience the maximum shielding effect of its π-electron cloud, with their signals observed in the strong field region, whereas the signals of the other protons located further from the macrocycle appear in the weak field region [67].
However, it is shown that in the presence of the anion, the complex formation is accompanied by the formation of an internal 1:1 complex (66) (Scheme 20) with the stability constant Ka = 6.8 × 106 M−1. The formation of the internal complexes is confirmed by the fact that the spectrum shows only one signal of the substrate protons located in the strong field region. The same shielding of the substrate protons by the receptor porphyrin fragments leads to the appearance of signals of the equivalent protons experiencing the shielding effect of the porphyrin fragment π-electron cloud in the strong field region. Thus, the conformational changes caused by the formation of the calix[4]pyrrole fragment complex with the anion make the receptor porphyrin fragments move closer to each other and facilitate the formation of internal receptor–substrate complexes with two points of binding.
A supramolecular electron donor–acceptor ensemble composed of three different components was created on the basis of a porphyrin anion, radical cation of tetrathiafulvalene calix[4]pyrrole (67), and radical anion of Li+-encapsulated C60 (Li+-C60) (Figure 10) [9].
These components bind to one another in a specific fashion via a combination of electrostatic and donor–acceptor interactions (Figure 11). It was shown that binding of a porphyrin carboxylate anion to tetrathiafulvalenecalix[4]pyrrole results in electron transfer from 67 to Li+-C60 to produce the charge-separated state (68) in benzonitrile.
Upon photoexcitation of (66), photoinduced electron transfer from the triplet excited state of porphyrin carboxylate to 67 occurs to produce the higher energy charge-separated state (68), which decays to the ground state with a lifetime of 4.8 µs, as determined from the absorption recovery of the bleaching band at 650 nm ascribed to the porphyrin ground state (Figure 12). This study provides a new approach to creating noncovalent, multi-component systems that support the formation of photoinduced charge-separated states.
In work [44], the authors synthesized oxacalix-[2]-arene-[2]-Bis-(Zn-porphyrin) (69) through nucleophilic substitution of the AB3-Zn-porphyrin hydroxy-derivative (70) for 5,17-Bis-(methylsulfonyl)-oxacalixarene (71) (Scheme 21).
The 1H-NMR titration method showed that receptor 69 effectively binds the C70 fullerene molecules in a solution [44]. At the same time, the authors note that this receptor does not interact with the C60 fullerene. It means that if the C70 fullerene molecule fits into the interporphyrin complexing cavity of 69, it creates the ideal conditions for substrate–receptor interactions leading to the formation of s Table 1:1 (69-C70) complexes (Figure 13) with the stability constant Ka = 30,000 M−1.
The authors of [77] synthesized a stable supramolecular capsule (72) based on tetracarboxy-resorcinarene (73) and 2-pyridylporphyrin (74) (Scheme 22). The 1H-NMR method showed that receptor 72 can effectively bind small organic molecules of various sizes. Figure 14A shows the spectra of the porphyrin (a), resorcinarene (b), and supramolecular capsule on their basis (c). The spectrum (Figure 14A(c)) clearly shows that the formation of supramolecular complex 72 shifts the signal of the peripheral resorcinarene protons up-field. The fact that the biggest shift of the signals of the protons (Hi) is observed in the resorcinarene -OCH2O- fragments located closer to the porphyrin macrocycle indicates that the resorcinarene fragment is located above the tetrapyrrole macrocycle.
Figure 14A(d) shows the 1H-NMR spectrum of the capsule 72 complex with an ethylene molecule. The strong field signal (in the region of −0.80 ppm) appearing in the spectrum when gaseous ethylene is purged through the receptor solution indicates the formation of an internal ethylene complex with the porphyrin capsule. It is shown (Figure 14B) that capsule 72 can selectively bind methane, ethane, propane, and ethylene molecules. The complex stability value (M−1, at 298 K) changes depending on the substrate molecule’s nature as follows: methane (56) < acetone (202) < ethylene (690) > ethane (527) > propane (236) > 2-methylpropane (3).

3. Molecular Receptors Based on Porphyrins Modified at the Macrocycle Periphery with Bulky Substituents

An important direction in the design of tetrapyrrole macrocyclic receptors for a certain substrate type is modification of the macrocycle periphery with bulky substituents or molecular fragments of different natures [81,82,83,84,85,86,87,88,89,90,91,92]. Bulky highly-branched lateral substituents are capable of forming additional complexing cavities that can be used for identification and selective binding of substrates of a certain type [93,94,95,96,97,98,99,100,101,102,103,104,105].
The authors of [103] obtained Zn-complexes of 5,15-diphenylporphyrins with alkoxy-substituents of various lengths in the macrocycle phenyl fragments (7577) through condensation of dipyrromethane (78) with the respective aromatic aldehydes in the presence of zinc acetate and catalytic amounts of an acid, followed by oxidation of the intermediate porphyrinogens by ortho-chloranil (Scheme 23).
It is shown that porphyrinates form 1:1 complexes with imidazole (Im), propylamine (PrA), 1,3-diaminopropane (DAPr), and DABCO (Figure 15). The authors attribute the higher stability of complexes of dimethoxy-substituted porphyrinate 76 with propylamine (76-PrA, Ka = 25200 M−1) than that of the similar complex of diphenylporphyrinate 75 (75-PrA, Ka = 17800 M−1) to the presence of ortho-methoxy groups in its aryl fragments.
Despite the existence of two isomers of porphyrinate 76 with different positions of the alkoxy-aryl fragments relative to the tetrapyrrole macrocycle, both isomers can form receptor–substrate complexes with two binding sites (between the macrocycle zinc cation and the ligand nitrogen atom and between the macrocycle oxygen atom and the ligand hydrogen one), in contrast to diarylporphyrinate with a single-site interaction between the receptor molecule and the substrate (Figure 16).
The bigger increase in the stability constants of tetradodecyloxy-substituted porphyrinate 77 with DAPr (77-DAPr, Ka = 54200 M−1) than that in the respective complex of dimethoxy-substituted porphyrinate 76 (76-DAPr, Ka = 27400 M−1) is explained by the presence of alkoxy-substituents at both sides of the macrocycle coordination center (Figure 17). If the porphyrinate and substrate reaction centers match geometrically, complexes with three binding points are formed between the zinc dodecyl-substituted porphyrinate and Bis-amine [103]. The structure of the receptor–substrate complex was verified by the 2D ROESY NMR method.
It should be said that a lot of works describe the binding ability of macrocyclic receptors with well-defined geometries [105,106,107,108,109,110,111]. An analysis of the presented data indicates that their complexing ability is strongly dependent on the conformational mobility of the lateral substituents (molecular fragments) and the possibility of formation of intramolecular cavities and channels of various shapes within them.
The authors of [109] synthesized zinc porphyrinate (79) with first-generation carbazole branches by the reaction of azide-alkyne cycloaddition of meso-ethynylphenyl substituted carbazole (80) with the octa-azido derivative of zinc porphyrinate (81) in the presence of copper chloride (Scheme 24). In a similar way, they synthesized zinc porphyrinate (82) with peripheral dendrimer second-generation branching based on 81 and meso-ethynylphenyl substituted second-generation carbazole (83). Structures of the carazolylphenyl branches of the first- (79,80) and second- (82,83) generations are depicted in the Scheme 25.
It was established by the spectrophotometric titration and 1H-NMR methods that the formation of a complex between zinc porphyrinate (79) and triazole (TrA) (Scheme 26) leads to the formation of one family of spectral curves corresponding to its own set of isobestic points (Figure 18a).
The titration curve has one stage, which indicates the formation of one type of complexes of 1:1 composition (Figure 18b). The same results are obtained when TrA forms a complex with zinc porphyrinate with two and four carbazole branches (84,85) prepared in a similar way to 79 on the basis of di- and tetra-azido derivatives of the respective zinc porphyrinates.
In the case of di- and tetra-substituted porphyrinates (84,85) (Figure 19), the triazole-carbazole fragments are bound with the macrocycle by a benzene bridge, i.e., there are no bulky substituents in the ortho-position of the benzene bridge. Unsubstituted phenyl fragments in meso-tetraphenylporphin are known to possess sufficient conformational mobility and can rotate along the carbon–carbon bond linking the phenyl fragment with the tetrapyrrole macrocycle.
The stability constants of the respective complexes for zinc tetraphenylporphyrinate (86) and zinc porphyrinates with two (84), four (85) and eight (79) first-generation carbazole branches and the studied ligands are given in Table 6 and are shown in Figure 20 as a plot [109].
The data show that the most stable complexes are those formed with the porphyrinate with eight carbazole fragments (79). This is probably the result of the specific structure of the considered macrocyclic receptors.
Probably, in the case of di- and tetra-substituted structures, most of the carbazole branches are not fixed relative to each other but can quite freely rotate relative to the tetrapyrrole macrocycle. As a result, there are no clearly defined and spatially fixed complexing cavities in these porphyrinates. However, in the case of the octa-substituted porphyrinate (79), the aromatic branches are mostly fixed spatially due to the ortho-methyl groups in the benzene bridge linking the porphyrin and carbazole fragments and, probably, forming complexing cavities above and under the tetrapyrrole macrocycle plane. Due to the fact that the substrate fits well into the intramolecular cavity of the porphyrin receptor and that there are additional hydrogen bonds and/or π–π interactions between the ligand and the triazole fragments of the porphyrinate, the internal complex formed as a result is characterized by increased stability (Scheme 27).
Interesting results were obtained by the authors of [109] when studying the ability of carbazolyl-substituted zinc porphyrinates to form complexes with triethylenediamine, a two-centered substrate. Di- and tetra-substituted porphyrinates 84 and 85 were found to interact with DABCO, and the process consisted of two stages. The complex formation spectrum has two families of spectral curves (Figure 21), each with its own family of isobestic points. The logarithmic dependence of the changes in the optical density of the system on ligand concentration also has two stages (Figure 22).
The fact that the 1H-NMR spectrum of the complex formed in the region of the low ligand concentration has one signal of the bound DABCO protons (−3.4 ppm) indicates that the complex formed has a 2:1 composition and its stability constant Ka = 4.1 × 109 M−2. The sandwich structure of the complex obtained when the ligand is located between the same porphyrin macrocycles makes all the ligand protons equivalent and leads to their appearance in the form of one signal in the strong field region.
The splitting of the proton signals in the complex 1H-NMR spectrum at higher ligand concentrations, the second stage on the titration curve (Figure 22), indicates that the complex formed has a 1:1 composition with the stability constant Ka = 6.9 × 106 M−1. Some of the protons located near the macrocycle experience the maximum shielding effect of its π-electron system, and their signals are observed in the strong field region (−3.4 ppm). The shielding effect on the other protons located at a distance from the macrocycle is weaker, and their signals are observed in the weak field region (0.5 ppm). The porphyrinate with four branches interacts with DABCO in the same way but the complex formation occurs at higher ligand concentration. At the same time, it is shown that the complex formation of octa-substituted porphyrinate (79) with DABCO consists of only one stage. Evidently, the bulky spatially fixed substituents prevent two-point interaction with DABCO leading to the formation of 1:1 complexes only. The stability values of the 36-DABCO and 39-DABCO complexes are comparable.
The authors of [110] investigated the complex formation between carbazolyl-substituted zinc porphyrinates with first- (79) and second-generation (82) lateral branches and cryptands that are significantly different in their geometric and complexing parameters and the effect of alkali metal cations on this process. Cryptands are well known to be able to act as receptors of alkali metal cations. The binding selectivity is the result of the oxygen-containing three-dimensional complexing cavities in them allowing them to identify cations of different natures. Table 7 presents the binding constants of alkali metal cations by cryptand-[2,2,2] in methanol at 25 °C.
It should be said that complex formation is accompanied by considerable changes in the cryptand conformation, according to [111]. Figure 23 shows an example of complex formation between cryptand-[2,2,2] and a potassium cation and conformational changes accompanying this process in the complex.
On the other hand, cryptands themselves can act as substrates, for example, in the process of complex formation with metal porphyrinates. The cryptand nitrogen atoms and the porphyrinate metal cation act in this case as the binding centers. The authors of [110] studied the formation of a complex between zinc porphyrinate 82 and cryptand-2,2,2 (Cr1) as a model system. It is shown that the electronic absorption spectra of the reaction system have one family of spectral curves with its own set of isobesticpoints (Figure 24a). The titration curve has one stage showing that the complex formation process consists of one stage only (Figure 24b).
The complex stability constant value is shown in Table 8 and is generally lower than those of the zinc tetraphenylporphyrinate (82) complexes with nitrogen-containing bases.
As it has been mentioned above, the bulky carbazole substituents at the periphery of the tetrapyrrole macrocycle in octa-substituted zinc porphyrinate 79 lead to the formation of complexing cavities. A study of complex formation between the octa-substituted zinc porphyrinate with the cryptand showed that the stability of the 1:1 complexes (79-Cr1) that are formed as a result (Scheme 28) is much higher than that of the similar zinc tetraphenylporphyrinate complex with the cryptand (82-Cr1) (Table 6). Evidently, the intramolecular complexing cavity formed by the carbazole branches has a positive effect on the complex formation due to the good agreement between the substrate and receptor geometry.
The 1H-NMR data also indicate the formation of a 1:1 complex in the region of equivalent reagent concentrations (Figure 25). The spectrum of the initial ligand complex with the potassium cation has three signals of the substrate protons—the signals of the central (-OCH2CH2O-) and lateral (-OCH2- and -CH2N-) protons.
In the 79-Cr1 porphyrinate complex, the -OMH2- and -CH2N- proton signals split. The signals of some of the protons located closer to the macrocycle are observed in the strong field region. The others remain in the weak field.
The same authors compared the ability of first- (79) and second-generation (82) dendrimers to form complexes with cryptands with considerable differences in geometry (Cr1-Cr3, Figure 26), including the effect of the alkali metal cation (Li+, Na+, and K+) on the substrate–receptor interactions.
Figure 26 shows that the stability of the complex that has no peripheral zinc porphyrinate 86 substituents with Cr1-Cr3 cryptands is much lower than that of the respective complexes of dendrimers 79 and 82.
In this case, dendrimer 82 with second-generation branches effectively binds the Cr1 and Cr2 ligands but does not interact with the Cr3 one, which is the largest in size. Cr3 is, evidently, unable to penetrate into the complexing cavity of the second-generation dendrimer because of its size.
A spectrophotometric study of the interaction between carbazolyl-substituted zinc porphyrinate 79 and Cr1 in the presence of alkali metal cations was conducted in a binary toluene–methanol solvent (9:1). The methanol additions are used to increase the solubility of lithium, sodium, and potassium bromides. Moreover, it should be taken into account that the alcohol additions weaken the coordination interaction between the porphyrinate zinc cation and the cryptand nitrogen atom. As a result, the stability of the receptor–substrate complexes formed becomes lower. It has been established that potassium bromide additions to the 79-Cr1 complex solution destroy the guest–host interaction (Scheme 29).
This hypothesis is, first of all, confirmed by the blue shift of the Soret band on the electronic absorption spectrum of the porphyrinate–cryptand reaction system (Figure 27a), indicating that the porphyrinate zinc cation does not interact with the cryptand nitrogen atom.
Secondly, the fact that the -OMH2- and -NCH2- proton signals in the NMR spectrum of the reaction system are not split (Figure 27b) also confirms the porphyrinate–cryptand complex dissociation in the presence of the alkali metal cation.
As Figure 28a shows, the presence of the lithium and sodium cations has practically no effect on the complex formation process considered. At the same time, the destruction of the 79-Cr2 complexes can only occur in the presence of lithium cations (Figure 28b).
Thus, complex formation inside the complexing cavity of the cryptand affects the formation of a complex between the carbazolyl-substituted zinc porphyrinate and the cryptand. The obtained experimental data can be explained by the effect of two factors. Firstly, as the calculations confirm, the formation of a complex between a cryptand and alkali metal cations is accompanied by changes in the molecule conformation as a whole. Secondly, the cryptands in the complexes with alkali metal cations are in the endo-endo conformation, in which the lone electron pairs of the nitrogen atoms face the ligand complexing cavity and, as a result, are inaccessible for coordination interaction with the porphyrinate zinc cation. The total effect of these factors, on the one hand, leads to the destruction of the porphyrinate–cryptand complex following the addition of a metal cation to the solution and, vice versa, prevents the zinc porphyrinate interaction with the cryptand of the corresponding metal cation.
It should be also said that the 79-Cr1 complex is destroyed as well when trifluoroacetic acid is added to the solution (Scheme 30). If the porphyrinate–ligand complex destruction under the action of the metal cation is caused by changes in the cryptate geometry in comparison with that of the cryptand, the complex destruction under the action of the acid is, evidently, caused by the breakage of the zinc–nitrogen bond between the porphyrinate metal cation and cryptand nitrogen upon protonation of the cryptand nitrogen atom.
The authors of [112] synthesized two-layered dendrimers (87G1 and 87G2) with inner carbazole and outer phenylazomethine layers by carrying out a reaction of zinc tetra(4-aminophenyl)porphyrinate (87) with the respective dendrons (G1 and G2) in chlorobenzene in the presence of TiCl4 (Scheme 31).
The 87G1 and 87G2 receptors have binding centers of two types: (1) a complexing cavity formed by carbazole fragments of the dendrimer outer layer and (2) a zinc cation of the tetrapyrrole macrocycle. It was established that 87G1 and 87G2 exhibited selective complexing ability towards the C60, C70, and C84 fullerenes. The binding represents a 1:1 type of interaction with the formation of complexes, in which one fullerene molecule is located in the complexing cavity made up of the carbazole fragments of the receptor outer layer. The stability constants of the respective complexes are given in Table 9. It has been established that in the case of 87G2, the binding capacity of C60 depends on the presence of phenyl-pyridine (PhPy) in the reaction mixture (the stability constant of the [87G2-C60]-PhPy complex is 1.5 times higher than the respective value of the [87G2-C60] complex).
According to [113], the formation of a complex between the porphyrinate zinc and the phenyl-pyridine nitrogen atom (Figure 29) causing conformational changes in the macrocycle dendrimer environment affects the ability of the outer carbazole layer of the 87G2 receptor to form complexes with C60.

4. Conclusions

The presented results show that porphyrin-containing macrocyclic compounds with a predetermined architecture and functional properties can be used to design ion-controlled molecular devices with a broad spectrum of action. The selectivity and high sensitivity of porphyrins to low-energy external effects make it possible to control chemical processes involving them. These characteristics together with the easily tunable shapes of molecules make the considered porphyrin-based molecular assemblies ideal receptors for extraction, directional transport, and sustained release of substances. Due to their extremely high extinction coefficients, porphyrins are also excellent molecular detectors for identification of processes affecting the tetrapyrrole macrocycle itself or the architecture surrounding it.

Author Contributions

Conceptualization and methodology, N.M. and O.K.; investigation, G.M. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Ministry of Science and Higher Education of the Russian Federation (№ 075-15-2021-579, in the part of studies of selective binding of cations and anions by tetrapyrrolic receptors), and the Russian Science Foundation (project № 19-73-20079, in the part of studies of selective binding of small organic molecules).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

This research was performed with the use of equipment of the Shared Facilities Center of the Institute of Organic Chemistry of the Russian Academy of Sciences and the Shared Facility Center, the Upper Volga Regional Center of Physicochemical Studies.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Beletskaya, I.; Tyurin, V.S.; Tsivadze, A.Y.; Gulard, R.; Stern, C. Supramolecular Chemistry of Metalloporphyrins. Chem. Rev. 2009, 109, 1659–1713. [Google Scholar] [CrossRef]
  2. Guchhait, T.; Sasmal, S.; Khan, F.S.T.; Rath, S.P. Oxo- and hydroxo-bridged diiron(III) porphyrin dimers: Inorganic and bio-inorganic perspectives and effects of intermacrocyclic interactions. Coord. Chem. Rev. 2017, 337, 112–144. [Google Scholar] [CrossRef]
  3. Mamardashvili, G.M.; Kaigorodova, E.Y.; Dmitrieva, O.; Mamardashvili, N.Z.; Koifman, O.I. Molecular recognition of imidazole derivatives by Co(III)-porphyrins in phosphate buffer (pH = 7.4) and cetylpyridinium chloride containing solutions. Molecules 2021, 26, 868. [Google Scholar] [CrossRef]
  4. Mamardashvili, N.Z.; Koifman, M.O. Synthesis of Monohydroxy-Substituted Diarylporphyrins and Their Binding Ability towards Aminobenzoic Acids. Macroheterocycles 2011, 4, 30–33. [Google Scholar] [CrossRef] [Green Version]
  5. Gil-Ramirez, G.; Karlen, S.D.; Shundo, A.; Porfyrakis, K.; Ito, Y.; Briggs, G.A.D.; Morton, J.J.; Anderson, H.L. A Cyclic Porphyrin Trimer as a Receptor for Fullerenes. Org. Lett. 2010, 12, 3544–3547. [Google Scholar] [CrossRef]
  6. Hu, J.; Liu, S. Engineering Responsive Polymer Building Blocks with Host–Guest Molecular Recognition for Functional Applications. Acc. Chem. Res. 2014, 47, 2084–2095. [Google Scholar] [CrossRef]
  7. Chandramouli, N.; Ferrand, Y.; Lautrette, G.; Kauffmann, B.; Mackereth, C.D.; Laguerre, M.; Dubreuil, D.; Huc, I. Iterative design of a helically folded aromatic oligoamide sequence for the selective encapsulation of fructose. Nat. Chem. 2015, 7, 334–341. [Google Scholar] [CrossRef] [PubMed]
  8. Margetić, D. Host-guest Studies of Bis-porphyrins. Curr. Org. Chem. 2012, 16, 829–851. [Google Scholar] [CrossRef]
  9. Davis, C.M.; Ohkubo, K.; Lammer, A.D.; Kim, D.S.; Kawashima, Y.; Sessler, J.L.; Fukuzumi, S. Photoinduced electron transfer in a supramolecular triad produced by porphyrin anion-induced electron transfer from tetrathiafulvalene calix[4]pyrrole to Li+@C60. Chem. Commun. 2015, 51, 9789–9792. [Google Scholar] [CrossRef] [Green Version]
  10. Liu, Q.; Dong, J.; Sun, Q.; Zhao, S.; Chen, Y.; Jiang, J. A novel calix[4]arene-modified porphyrin-baseddual-mode sensor for the specific detection ofdopamine with excellent performance. New J. Chem. 2019, 43, 10376–10381. [Google Scholar] [CrossRef]
  11. Nguyen, N.T.; Hofkens, J.; Scheblykin, I.G.; Kruk, M.; Dehaen, W. Click Reaction Synthesis and Photophysical Studies of Dendritic Metalloporphyrins. Eur. J. Org. Chem. 2014, 2014, 1766–1777. [Google Scholar] [CrossRef]
  12. Merkas, S.; Bouatra, S.; Rein, R.; Piantanida, I.; Zinic, M.; Solladié, N. Pre-organized dinucleosides with pendant porphyrins for the formation of sandwich type complexes with DABCO with high association constants. J. Porphyr. Phthalocyanines 2015, 19, 535–546. [Google Scholar] [CrossRef]
  13. Higashino, T.; Fujimori, Y.; Sugiura, K.; Tsuji, Y.; Ito, S.; Imahori, H. Synthesis of push–pull porphyrin with two electron-donating and two electron-withdrawing groups and its application to dye-sensitized solar cell. J. Porphyr. Phthalocyanines 2015, 19, 140–149. [Google Scholar] [CrossRef]
  14. Sharma, G.D.; Zervaki, G.E.; Ladomenou, L.; Koukaras, E.N.; Panagiotis, P.; Angaridis, P.P.; Coutsolelos, A.G. Donor-π-acceptor, triazine-linked porphyrin dyads as sensitizers for dye-sensitized solar cells. J. Porphyr. Phthalocyanines 2015, 19, 175–191. [Google Scholar] [CrossRef]
  15. Dhamija, A.; Mondal, P.; Saha, B.; Rath, S.P. Induction, control, and rationalization of supramolecular chirogenesis using metalloporphyrin tweezers: A structure-function correlation. Dalton Trans. 2020, 49, 10679–10700. [Google Scholar] [CrossRef]
  16. Milanesio, M.E.; Alvarez, M.G.; Durantini, E.N. Methoxyphenyl Porphyrin Derivatives as Phototherapeutic Agents. Curr. Bioact. Compd. 2010, 6, 97–105. [Google Scholar] [CrossRef]
  17. De Visser, S.P.; Valentine, J.S.; Nam, W. A Biomimetic Ferric Hydroperoxo Porphyrin Intermediate. Angew. Chem. Int. Ed. Engl. 2010, 49, 2099–2101. [Google Scholar] [CrossRef]
  18. Gonsalves, A.M.A.R.; Serra, A.C.; Pineiro, M. The small stones of Coimbra in the huge tetrapyrrolic chemistry building. J. Porphyr. Phthalocyanines 2009, 13, 429–445. [Google Scholar] [CrossRef]
  19. Maeda, C.; Kamada, T.; Aratani, N.; Osuka, A. Chiral self-discriminative self-assembling of meso-meso linked diporphyrins. Coord. Chem. Rev. 2007, 251, 2743–2752. [Google Scholar] [CrossRef]
  20. Hori, T.; Nakamura, Y.; Aratani, N.; Osuka, A. Exploration of electronically interactive cyclic porphyrin arrays. J. Organomet. Chem. 2007, 692, 148–155. [Google Scholar] [CrossRef]
  21. Mamardasvili, G.M.; Mamardashvili, N.Z.; Koifman, O.I. Self-assembling systems based on porphirins. Russ. Chem. Rev. 2008, 77, 765–782. [Google Scholar] [CrossRef]
  22. Mamardashvili, G.M.; Mamardashvili, N.Z.; Koifman, O.I. Synthesis and receptor properties of calix[4]pyrroles. Russ. Chem. Rev. 2015, 84, 275–287. [Google Scholar] [CrossRef]
  23. Balaban, T.S. Tailoring Porphyrins and Chlorins for Self-Assembly in Biomimetic Artificial Antenna Systems. Acc. Chem. Res. 2005, 38, 612–623. [Google Scholar] [CrossRef]
  24. Carofiglio, T.; Lubian, E.; Varotto, A. Synthesis, heterogenization and sensing properties of melamine-bridged Bis-porphyrin dimers. J. Porphyr. Phthalocyanines 2010, 14, 701–707. [Google Scholar] [CrossRef]
  25. Samaroo, D.; Vinodu, M.; Chen, X.; Drain, C.M. meso-Tetra(pentafluorophenyl)porphyrin as an Efficient Platform for Combinatorial Synthesis and the Selection of New Photodynamic Therapeutics using a Cancer Cell Line. J. Comb. Chem. 2007, 9, 998–1011. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  26. Tsuda, A. Design of Porphyrin Nanoclusters toward Discovery of Novel Properties and Functions Bull. Chem. Soc. Jpn. 2009, 82, 11–28. [Google Scholar] [CrossRef] [Green Version]
  27. Goldberg, I. Crystal engineering of nanoporous architectures and chiralporphyrinassemblies. CrystEngComm 2008, 10, 637–645. [Google Scholar] [CrossRef]
  28. Boyd, P.D.W.; Reed, C.A. Fullerene−Porphyrin Constructs. Acc. Chem. Res. 2005, 38, 235–242. [Google Scholar] [CrossRef] [Green Version]
  29. Jurow, M.; Schuckman, A.E.; Batteas, J.D.; Drain, C.M. Porphyrins as molecular electronic components of functional devices. Coord. Chem. Rev. 2010, 254, 2297–2310. [Google Scholar] [CrossRef] [Green Version]
  30. Yua, G.; Zhub, B.; Shaoc, L.; Zhouc, J.; Sahad, M.L.; Shid, B.; Zhang, Z.; Hong, T.; Lib, S.; Chena, X.; et al. Host−guest complexation-mediated codelivery of anticancer drug and photosensitizer for cancer photo-chemotherapy. Proc. Natl. Acad. Sci. USA 2019, 116, 6618–6623. [Google Scholar] [CrossRef] [Green Version]
  31. Vinodh, M.; Alipour, F.H.; Abdirahman, A.M.; Talal, F. Al-Azemi Molecular Assemblies of Porphyrins and Macrocyclic Receptors: Recent Developments in Their Synthesis and Applications. Molecules 2012, 17, 11763–11799. [Google Scholar] [CrossRef] [Green Version]
  32. Blom, M.; Norrehed, S.; Andersson, C.-H.; Huang, H.; Light, M.E.; Bergquist, J.; Grennberg, H.; Gogoll, A. Synthesis and Properties of Bis-Porphyrin Molecular Tweezers: Effects of Spacer Flexibility on Binding and Supramolecular Chirogenesis. Molecules 2016, 21, 16. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  33. Liang, X.; Honglin Zhang, H.; Luo, H.; Qin, M.; Li, M.; Zhu, W. Bio-Inspired Ni(II) Porphyrin Dimers with a Bridging Diphenyl Moiety: Facile Synthesis and Molecular Inherent Chirality. Macroheterocycles 2019, 12, 35–39. [Google Scholar] [CrossRef]
  34. Huang, X.; Borhan, B.; Rickman, B.H.; Nakanishi, K.; Berova, N. Binding of Cationic Bis-porphyrins Linked with p- or m- Xylylenediamine and Their Zinc(II) Complexes to Duplex DNA. Molecules 2008, 13, 3117–3128. [Google Scholar] [CrossRef] [Green Version]
  35. Ballester, P.; Costa, A.; Castilla, A.M.; Dey, P.M.; Frontera, A.; Gomila, R.M.; Hunter, C.A. DABCO-Directed Self-Assembly of Bisporphyrins (DABCO=1,4-Diazabicyclo[2.2.2]octane). Chem. Eur. J. 2005, 11, 2196. [Google Scholar] [CrossRef]
  36. Dudic, M.; Lhoták, P.; Proškova, P.; Stibor, I.; Lang, K.; Sýkora, J. Calixarene-based metalloporphyrins: Molecular tweezers for complexation of DABCO. Tetrahedron 2003, 59, 2409–2415. [Google Scholar] [CrossRef]
  37. Káš, M.; Lang, K.; Stibor, I.; Lhoták, P. Novel fullerene receptors based on calixarene-porphyrin conjugates. Tetrahedron Lett. 2007, 48, 477–481. [Google Scholar] [CrossRef]
  38. Hosseini, A.; Taylor, S.; Accorsi, G.; Armaroli, N.; Reed, C.A.; Boyd, P.D. Calix[4]arene-Linked Bisporphyrin Hosts for Fullerenes:  Binding Strength, Solvation Effects, and Porphyrin−Fullerene Charge Transfer Bands. J. Am. Chem. Soc. 2006, 128, 15903–15913. [Google Scholar] [CrossRef] [Green Version]
  39. Jokic, D.; Zouhair Asfari, Z.; Weiss, J. The First Versatile Synthetic Approach to Cofacial Bis-Porphyrins with Calixarene Spacers. Org. Lett. 2002, 4, 2129–2132. [Google Scholar] [CrossRef]
  40. Mamardashvili, G.M.; Shinkar’, I.A.; Mamardashvili, N.Z.; Koifman, O.I. Calix[4]arene-Porphyrin Molecular Receptors for Selective Binding of Ethylenediamines. Russ. J. Coord. Chem. 2007, 33, 774–778. [Google Scholar] [CrossRef]
  41. Mamardashvili, G.M.; Mamardashvili, N.Z.; Koifman, O.I. Cation-Dependent Binding of Zinc Diethoxycarbonylcalix[4]arene-Bis(porphyrinate) Triethylenediamine. Russ. J. Coord. Chem. 2011, 37, 195–201. [Google Scholar] [CrossRef]
  42. Dudic, M.; Lhoták, P.; Stibor, I.; Dvoráková, H. Lang Kamil Synthesis and spectroscopic properties of porphyrin-(thia)calix[4]arene conjugates. Tetrahedron 2002, 58, 5475–5482. [Google Scholar] [CrossRef]
  43. Kundrat, O.; Kas, M.; Tkadlecova, M.; Lang, K.; Cvacka, J.; Stibor, I.; Lhotak, P. Thiacalix[4]arene–porphyrin conjugates with high selectivity towards fullerene C70. Tetrahedron Lett. 2007, 48, 6620–6623. [Google Scholar] [CrossRef]
  44. Rossom, W.V.; Kundrát, O.; Ngo, T.H.; Lhoták, P.; Dehaen, W.; Maes, W. An oxacalix[2]arene[2]pyrimidine-Bis(Zn-porphyrin) tweezer as a selective receptor towards fullerene C70. Tetrahedron Lett. 2010, 51, 2423–2426. [Google Scholar] [CrossRef]
  45. Dudić, M.; Lhoták, P.; Stibor, I.; Lang, K.; Proškova, P. Calix[4]arene-porphyrin Conjugates as Versatile Molecular Receptors for Anions. Org. Lett. 2003, 5, 149–152. [Google Scholar] [CrossRef]
  46. Yakushev, A.A.; Averin, A.D.; Sakovich, M.V.; Vatsouro, I.M.; Kovalev, V.V.; Syrbu, S.A.; Koifman, O.I.; Beletskaya, I.P. Synthesisoftheporphyrin-calix[4]areneconjugates via Pd-catalyzedaminationandtheirevaluationasfluorescentchemosensors. J. Porphyr. 2019, 23, 1551–1562. [Google Scholar] [CrossRef]
  47. Iwamoto, H.; Nishi, S.; Haino, T. Highly shape-selective guest encapsulation in the precisely defined cavity of a calix[4]arene-capped metalloporphyrin. Chem. Commun. 2011, 47, 12670–12672. [Google Scholar] [CrossRef]
  48. Nagasaki, T.; Fujishima, H.; Shinkai, S. Calix[4]arene-Capped Tetraphenylporphyrin. Synthetic Approach to a Chiral Capped Porphyrin with Regular C4 Symmetry. Chem. Lett. 1994, 23, 989–992. [Google Scholar] [CrossRef]
  49. Iwamoto, H.; Yukimasa, Y.; Yoshimasa Fukazawa, Y. Synthesis and binding behavior of a Zn(II)-porphyrin having calix[5]arene cap. Tetrahedron Lett. 2002, 43, 8191–8194. [Google Scholar] [CrossRef]
  50. Mibradt, R.; Weiss, J. Efficient Combination of Calix[4]arenes and meso-Diphenylporphyrins. Tetrahedron Lett. 1995, 36, 2999–3002. [Google Scholar] [CrossRef]
  51. Pognon, G.; Boudon, C.; Schenk, K.J.; Bonin, M.; Bach, B.; Weiss, J. Electrochemically Triggered Open and Closed Pacman Bis-metalloporphyrins. J. Am. Chem. Soc. 2006, 128, 3488–3489. [Google Scholar] [CrossRef] [PubMed]
  52. Holler, M.; Schmitt, M.; Nierengarten, J.-F. Synthesis and conformational analysis of porphyrin derivatives substituted with calix[4]arene subunits. J. Porphyr. Phthalocyanines 2011, 15, 1183–1188. [Google Scholar] [CrossRef]
  53. Mamardashvili, N.Z.; Mamardashvili, G.M.; Weiss, J. Complexation of Zinc Porphyrintes with 1,4-Diazabicyclo[2,2,2]octane. Russ. J. Inorg. Chem. 2005, 50, 218–224. [Google Scholar]
  54. Mamardashvili, N.Z.; Koifman, O.I. Cation and anion assisted binding of triethylenediamine by zinc Bis-porphyrinates. Russ. Chem. Bull. 2013, 62, 123–132. [Google Scholar] [CrossRef]
  55. Puglisi, A.; Purrello, R.; Rizzarelli, E.; Sortino, S.; Vecchio, G. Spectroscopic and self-association behavior of a porphyrin-β-cyclodextrinconjugate. New J. Chem. 2007, 31, 1499–1506. [Google Scholar] [CrossRef]
  56. Hosokawa, K.; Miura, Y.; Kiba, T.; Kakuchi, T.; Sato, S. Fluorescence Resonance Energy Transfer in Host–Guest Inclusion Complexes of Cyclodextrin–Porphyrin Composite in Aqueous Solution. Chem. Lett. 2008, 37, 60–61. [Google Scholar] [CrossRef]
  57. Fathalla, M.; Li, S.; Diebold, U.; Alb, A.; Jayawickramarajah, J. Water-soluble nanorods self-assembledviapristine C60 and porphyrin moieties. Chem. Commun. 2009, 28, 4209–4211. [Google Scholar] [CrossRef]
  58. Kralova, J.; Kejik, Z.; Briza, T.; Pouckova, P.; Kral, A.; Martasek, P.; Kral, V. Porphyrin−Cyclodextrin Conjugates as a Nanosystem for Versatile Drug Delivery and Multimodal Cancer Therapy. J. Med. Chem. 2010, 53, 128–138. [Google Scholar] [CrossRef]
  59. Guo, Y.; Zhang, P.; Chao, J.; Shuang, S.; Dong, C. Study on the supramolecular system of 5-(p-hydroxyphenyl)-10,15,20-tris-(4-chlorophenyl)porphyrin with cyclodextrins and its analytical characteristics. Spectrochim. Acta A 2008, 71A, 946–950. [Google Scholar] [CrossRef]
  60. Fathalla, M.; Neuberger, A.; Li, S.-C.; Schmehl, R.; Diebold, U.; Jayawickramarajah, J. Straightforward Self-Assembly of Porphyrin Nanowires in Water: Harnessing Adamantane/β-Cyclodextrin Interactions. J. Am. Chem. Soc. 2010, 132, 9966–9967. [Google Scholar] [CrossRef] [PubMed]
  61. Guo, Y.-J.; Chao, J.-B.; Pan, J.-H. Study on the interaction of 5-pyridine-10,15,20-tris-(p-chlorophenyl)porphyrin with cyclodextrins and DNA by spectroscopy. Spectrochim. Acta A 2007, 68A, 231–236. [Google Scholar] [CrossRef]
  62. Kiba, T.; Suzuki, H.; Hosokawa, K.; Kobayashi, H.; Baba, S.; Kakuchi, T.; Sato, S. SupramolecularJ-Aggregate Assembly of a Covalently Linked Zinc Porphyrin−β-cyclodextrin Conjugate in a Water/Ethanol Binary Mixture. J. Phys. Chem. B 2009, 113, 11560–11563. [Google Scholar] [CrossRef] [PubMed]
  63. Ermilov, E.A.; Menting, R.; Lau, J.T.F.; Leng, X.; Roeder, B.; Ng, D.K.P. Switching the photoinduced processes in host–guest complexes of β-cyclodextrin-substituted silicon(iv) phthalocyanines and a tetrasulfonated porphyrin. Phys. Chem. Chem. Phys. 2011, 13, 17633–17641. [Google Scholar] [CrossRef]
  64. Xu, H.; Ermilov, E.A.; Roeder, B.; Ng, D.K.P. Formation and energy transfer property of a subphthalocyanine–porphyrin complex held by host–guest interactions. Phys. Chem. Chem. Phys. 2010, 12, 7366–7370. [Google Scholar] [CrossRef]
  65. Khavasi, H.R.; Sasan, K.; Safari, N. Inclusion complex of ironmeso-tetrakis(p-sulfonatophenyl)porphyrin and 2-hydroxypropyl-β-cyclodextrin as a functional model of cytochrome P-450: Study on a supramolecular formation and its application in aqueous oxidation of styrene. J. Porphyr. Phthalocyanines 2007, 11, 874–882. [Google Scholar] [CrossRef]
  66. Liu, Y.; Ke, C.-F.; Zhang, H.-Y.; Cui, J.; Ding, F. Complexation-Induced Transition of Nanorod to Network Aggregates:  Alternate Porphyrin and Cyclodextrin Arrays. J. Am. Chem. Soc. 2008, 130, 600–605. [Google Scholar] [CrossRef]
  67. Mamardashvili, G.M.; Mamardashvili, N.Z.; Koifman, O.I. Anion-Dependent Binding of Zinc Calix[4]pyrrole-Bisporphyrinate Triethylenediamine. Russ. J. Coord. Chem. 2011, 37, 872–877. [Google Scholar] [CrossRef]
  68. Donato Monti, D.; Monica, L.L.; Anita Scipioni, A.; Mancini, G. Effect of the inclusion of sodium cations on the binding properties of a switchable diporphyrin receptor. New J. Chem. 2001, 25, 780–782. [Google Scholar] [CrossRef]
  69. Iengo, E.; Zangrando, E.; Alessio, E.; Chambron, J.-C.; Heitz, V.; Flamigni, L.; Sauvage, J.-P. A Functionalized Noncovalent Macrocyclic Multiporphyrin Assembly from a Dizinc(II) Bis-Porphyrin Receptor and a Free-Base Dipyridylporphyrin. Chem. Eur. J. 2003, 9, 5879. [Google Scholar] [CrossRef]
  70. Flamigni, L.; Talarico, A.M.; Ventura, B.; Rein, R.; Solladié, N. A Versatile Bis-Porphyrin Tweezer Host for the Assembly of Noncovalent Photoactive Architectures: A Photophysical Characterization of the Tweezers and Their Association with Porphyrins and Other Guests. Chem. Eur. J. 2006, 12, 701. [Google Scholar] [CrossRef]
  71. Guldi, D.M.; Luo, C.; Swartz, A.; Scheloske, M.; Hirsch, A. Selforganisation in photoactive fullerene porphyrin based donor–acceptor ensembles. Chem. Commun. 2001, 1066–1067. [Google Scholar] [CrossRef]
  72. El-Khouly, M.E.; Hasegawa, J.; Momotake, A.; Sasaki, M.; Araki, Y.; Ito, O.; Arai, T. Intramolecular photoinduced processes of newly synthesized dual zinc porphyrin-fullerene triad with flexible linkers. J. Porph. Phthaloc. 2006, 10, 1380–1391. [Google Scholar] [CrossRef]
  73. Tashiro, K.; Aida, T. Metalloporphyrin hosts for supramolecular chemistry of fullerens. Chem. Soc. Rev. 2007, 36, 189–197. [Google Scholar] [CrossRef] [PubMed]
  74. D’Souza, F.; Gadde, S.; Zandler, M.E.; Itou, M.; Araki, Y.; Ito, O. Supramolecular complex composed of a covalently linked zinc porphyrin dimer and fulleropyrrolidine bearing two axially coordinating pyridine entities. Chem. Commun. 2004, 2276–2277. [Google Scholar] [CrossRef]
  75. Kubo, Y.; Ohno, T.; Yamanaka, J.; Tokita, S.; Iida, T.; Ishimaru, Y. Chirality-Transfer Control Using a Heterotopic Zinc(II) Porphyrin Dimer. J. Am. Chem. Soc. 2001, 123, 12700. [Google Scholar] [CrossRef]
  76. Brahma, S.; Ikbal, A.S.; Dhamija, A.; Rath, S.P. Highly Enhanced Bisignate Circular Dichroism of Ferrocene-Bridged Zn(II) Bisporphyrin Tweezer with Extended Chiral Substrates due to Well-Matched Host–Guest System. Inorg. Chem. 2014, 53, 2381–2395. [Google Scholar] [CrossRef]
  77. Nakazawa, J.; Mizuki, M.; Shimazaki, Y.; Tani, F.; Naruta, Y. Encapsulation of Small Molecules by a Cavitand Porphyrin Self-Assembled via Quadruple Hydrogen Bonds. Org. Lett. 2006, 8, 4275–4278. [Google Scholar] [CrossRef]
  78. Churakhina, Y.I.; Ivanova, Y.B.; Maltseva, O.V.; Mamardashvili, N.Z. Basic properties of porphyrins with polyethylenoxide spacers of various length. Russ. J. Gen. Chem. 2009, 79, 2435–2439. [Google Scholar] [CrossRef]
  79. Ho, K.-H.L.; Hijazi, I.; Rivier, L.; Gautier, C.; Jousselme, B.; de Miguel, G.; Romero-Nieto, C.; Guldi, D.M.; Heinrich, B.; Donnio, B.; et al. Host–Guest Complexation of [60]Fullerenes and Porphyrins Enabled by “Click Chemistry”. Chem. Eur. J. 2013, 19, 11374–11381. [Google Scholar] [CrossRef]
  80. Al-Azemi, T.F.; Vinodh, M. Synthesis of porphyrin conjugates based on conformationally rigid and flexible resorcin[4]arene frameworks. Tetrahedron 2011, 67, 2585–2590. [Google Scholar] [CrossRef]
  81. Shinoda, S.; Ohashi, M.; Tsukube, H. “Pocket Dendrimers” as Nanoscale Receptors for Bimolecular Guest Accommodation. Chem. Eur. J. 2007, 13, 81–89. [Google Scholar] [CrossRef] [PubMed]
  82. Shinoda, S. Nanoscale substrate recognition by porphyrin dendrimers with patched structures. J. Incl. Phenom. Macrocycl. Chem. 2007, 59, 1–9. [Google Scholar] [CrossRef]
  83. Huang, X.; Zhu, C.; Zhang, S.; Li, W.; Guo, Y.; Zhan, X.; Liu, Y.; Bo, Z. Porphyrin−Dithienothiophene π-Conjugated Copolymers: Synthesis and Their Applications in Field-Effect Transistors and Solar Cells. Macromolecules 2008, 41, 6895–6902. [Google Scholar] [CrossRef]
  84. Lee, C.Y.; Hupp, J.T. Dye Sensitized Solar Cells: TiO2 Sensitization with a Bodipy-Porphyrin Antenna System. Langmuir 2010, 26, 3760–3765. [Google Scholar] [CrossRef]
  85. Kruk, M.M.; Starukhin, A.S.; Mamardashvili, N.Z.; Mamardashvili, G.M.; Ivanova, Y.B.; Maltseva, O.V. Tetrapyrrolic compounds as hosts for binding of halides and alkali metal cations. J. Porphyr. Phthalocyanines 2009, 13, 1148–1158. [Google Scholar] [CrossRef]
  86. Li, L.; Huang, Y.; Peng, J.; Cao, Y.; Peng, X. Enhanced performance of solution-processed solar cells based onporphyrinsmall molecules with a diketopyrrolopyrrole acceptor unit and apyridineadditive. J. Mater. Chem. A 2013, 1, 2144–2150. [Google Scholar] [CrossRef]
  87. Favereau, L.; Warnan, J.; Anne, F.B.; Pellegrin, Y.; Blart, E.; Jacquemin, D.; Odobel, F. Diketopyrrolopyrrole-zinc porphyrin, a tuned panchromatic association fordye-sensitized solar cells. J. Mater. Chem. A 2013, 1, 7572–7575. [Google Scholar] [CrossRef]
  88. Maufroy, A.; Favereau, L.; Anne, F.B.; Pellegrin, Y.; Blart, E.; Hissler, M.; Jacquemin, D.; Odobel, F. Synthesis and properties of push–pull porphyrins as sensitizers for NiO based dye-sensitized solar cells. J. Mater. Chem. A 2015, 3, 3908–3917. [Google Scholar] [CrossRef] [Green Version]
  89. Higashino, T.; Imahori, H. Porphyrins as excellent dyes for dye-sensitized solar cells: Recent developments and insights. Dalton Trans. 2015, 44, 448–463. [Google Scholar] [CrossRef] [PubMed]
  90. Ballester, P.; Claudel, M.; Durot, S.; Kocher, L.; Choepffl, S.; Heitz, V. A Porphyrin Coordination Cage Assembled from Four Silver(I) Triazolyl-Pyridine Complexes. Chem. Eur. J. 2015, 21, 15339–15348. [Google Scholar] [CrossRef]
  91. Durot, S.; Taesch, J.; Heitz, V. Multiporphyrinic Cages: Architectures and Functions. Chem. Rev. 2014, 114, 8542–8578. [Google Scholar] [CrossRef] [PubMed]
  92. Qiu, W.-G.; Li, Z.-F.; Bai, G.-M.; Meng, S.-N.; Dai, H.-X.; He, H. Study on the inclusion behavior betweenmeso-tetrakis[4-(3-pyridiniumpropoxy)phenyl]prophyrin tetrakisbromide and β-cyclodextrin derivatives in aqueous solution. Spectrochim. Acta A 2007, 66A, 1189–1193. [Google Scholar] [CrossRef]
  93. Guo, Y.-J.; Guo, L.; Pan, J.-H. Study on the supramolecular system of tetrakis(2-hydroxy-5-sulfonatophenyl)porphyrin with cyclodextrins by spectroscopy. Phys. Chem. Liq. 2007, 45, 261–269. [Google Scholar] [CrossRef]
  94. Tsuchiya, Y.; Yamano, A.; Shiraki, T.; Sada, K.; Shinkai, S. Single-crystal Structure of Porphyrin Bicapped with Trimethyl-β-cyclodextrins: A Novel Dye-oriented Material. Chem. Lett. 2011, 40, 99–101. [Google Scholar] [CrossRef]
  95. Callari, F.L.; Mazzaglia, A.; Scolaro, L.M.; Valli, L.; Sortino, S. Biocompatible nanoparticles of amphiphilic cyclodextrins entangling porphyrins as suitable vessels for light-induced energy andelectron transfer. J. Mater. Chem. 2008, 18, 802–805. [Google Scholar] [CrossRef]
  96. Yu, M.; Chen, Y.; Zhang, N.; Liu, Y. Construction and transmembrane dissociation behavior of supramolecular assembly of quinolinocyclodextrin with porphyrin. Org. Biomol. Chem. 2010, 8, 4148–4154. [Google Scholar] [CrossRef]
  97. Ferro, S.; Jori, G.; Sortino, S.; Stancanelli, R.; Nikolov, P.; Tognon, G.; Ricchelli, F.; Mazzaglia, A. Inclusion of 5-[4-(1-Dodecanoylpyridinium)]-10,15,20-triphenylporphine in Supramolecular Aggregates of Cationic Amphiphilic Cyclodextrins: Physicochemical Characterization of the Complexes and Strengthening of the Antimicrobial Photosensitizing Activity. Biomacromolecules 2009, 10, 2592–2600. [Google Scholar] [CrossRef]
  98. Mora, S.J.; Cormick, M.P.; Milanesio, M.E.; Durantini, E.N. The photodynamic activity of a novel porphyrin derivative bearing a fluconazole structure in different media and against Candida albicans. Dye. Pigm. 2010, 87, 234–240. [Google Scholar] [CrossRef]
  99. Zhang, Y.-M.; Chen, Y.; Zhuang, R.-J.; Liu, Y. Supramolecular architecture of tetrathiafulvalene-bridged Bis(β-cyclodextrin) with porphyrin and its electron transfer behaviors. Photochem. Photobiol. Sci. 2011, 10, 1393–1398. [Google Scholar] [CrossRef]
  100. Zhang, Y.-M.; Chen, Y.; Yang, Y.; Liu, P.; Liu, Y. Supramolecular Architectures by Fullerene-Bridged Bis(permethyl-β-cyclodextrin)s with Porphyrins. Chem. Eur. J. 2009, 15, 11333–11340. [Google Scholar] [CrossRef]
  101. Valli, L.; Giancane, G.; Mazzaglia, A.; Scolaro, L.M.; Conoci, S.; Sortino, S. Photoresponsive multilayer films by assembling cationic amphiphiliccyclodextrinsandanionicporphyrinsat the air/waterinterface. J. Mater. Chem. 2007, 17, 1660–1663. [Google Scholar] [CrossRef]
  102. Wang, K.-R.; Guo, D.-S.; Jiang, B.-P.; Liu, Y. Excitonic coupling interactions in the self-assembly of perylene-bridged Bis-(β-cyclodextrin)s and porphyrin. Chem. Commun. 2012, 48, 3644–3646. [Google Scholar] [CrossRef] [PubMed]
  103. Maltceva, O.; Mamardashvili, G.; Khodov, I.; Lazovskiy, D.; Krest’yaninov, M.; Mamardashvili, N.; Dehaen, W. Molecular recognition of nitrogen—Containing bases by Zn[5,15-Bis-(2,6-dodecyloxyphenyl)]porphyrin. Supramol. Chem. 2017, 29, 360–369. [Google Scholar] [CrossRef]
  104. Long, L.; Jin, J.Y.; Zhang, Y.; Yang, R.; Wang, K. I sensing approach for albumin nteractionsbetweenproteinand porphyrin-containingcyclodextrinsupramolecular system: A fluorescent. Analyst 2008, 133, 1201–1208. [Google Scholar] [CrossRef]
  105. Oliveri, V.; Puglisi, A.; Vecchio, G. New conjugates of β-cyclodextrin with manganese(iii) salophen and porphyrin complexes as antioxidant systems. Dalton Trans. 2011, 40, 2913–2919. [Google Scholar] [CrossRef]
  106. Kano, K. Porphyrin-cyclodextrin supramolecular complexes as myoglobin model in water. Colloid Polym. Sci. 2008, 286, 79–84. [Google Scholar] [CrossRef]
  107. Watanabe, K.; Kitagishi, H.; Kano, K. Supramolecular Ferric Porphyrins as Cyanide Receptors in Aqueous Solution. ACS Med. Chem. Lett. 2011, 2, 943–947. [Google Scholar] [CrossRef] [Green Version]
  108. Kano, K.; Kitagishi, H.; Dagallier, C.; Kodera, M.; Matsuo, T.; Hayashi, T.; Hisaeda, Y.; Hirota, S. Iron Porphyrin−Cyclodextrin Supramolecular Complex as a Functional Model of Myoglobin in Aqueous Solution. Inorg. Chem. 2006, 45, 4448–4460. [Google Scholar] [CrossRef]
  109. Nguyen, N.T.; Mamardashvili, G.M.; Kulikova, O.M.; Scheblykin, I.G.; Mamardashvili, N.Z.; Dehaen, W. Binding ability of first and second generation/ carbazolylphenyl dendrimers with Zn(II) tetraphenylporphyrin core towards small heterocyclic substrates. RSC Adv. 2014, 4, 19703–19709. [Google Scholar] [CrossRef]
  110. Mamardashvili, G.M.; Maltceva, O.V.; Mamardashvili, N.Z.; Nguyen, N.T.; Dehaen, W. Cation assisted complexation of octacarbazolylphenyl substituted Zn(II)-tetraphenylporphyrin with [2,2,2]cryptand. RSC Adv. 2015, 5, 44557–44562. [Google Scholar] [CrossRef]
  111. Bukhzam, A.; Bade, N. Crown Ethers: Their Complexes and Analytical Applications. J. Applic. Chem. 2014, 3, 237–244. [Google Scholar]
  112. Albrecht, K.; Kasai, Y.; Kimoto, A.; Yamamoto, K. The Synthesis and Properties of Carbazole−Phenylazomethine Double Layer-Type Dendrimers. Macromolecules 2008, 41, 3793–3800. [Google Scholar] [CrossRef]
  113. Albrecht, K.; Kasai, Y.; Kuramoto, Y.; Yamamoto, K. Dynamic control of dendrimer–fullereneassociation by axial coordination to the core. Chem. Com. 2013, 46, 6861–6863. [Google Scholar] [CrossRef]
Scheme 1. (a) Pyrrole, TFA, 40%; (b) benzaldehyde, DDQ, CH2Cl2, 17%; (c) LiOH, THF, 92%; (d) (COCl)2, CH2Cl2; (e) Et3N, CH2Cl2, 22%; (f) Zn(OAc)2, MeOH, CHCl3, 74%.
Scheme 1. (a) Pyrrole, TFA, 40%; (b) benzaldehyde, DDQ, CH2Cl2, 17%; (c) LiOH, THF, 92%; (d) (COCl)2, CH2Cl2; (e) Et3N, CH2Cl2, 22%; (f) Zn(OAc)2, MeOH, CHCl3, 74%.
Molecules 26 05292 sch001
Figure 1. 1H-NMR spectra of the 7-Py (a) and 7-Im (b) complexes in CDCl3 at 298 K. The o and * symbols denote the signals of the free and bound substrate molecules [47].
Figure 1. 1H-NMR spectra of the 7-Py (a) and 7-Im (b) complexes in CDCl3 at 298 K. The o and * symbols denote the signals of the free and bound substrate molecules [47].
Molecules 26 05292 g001
Figure 2. Structure of the 7-Py complex determined by the molecular mechanics methods using the MMFF force field of the MacroModel (V6.5) software package on the SGI O2 workstation [56].
Figure 2. Structure of the 7-Py complex determined by the molecular mechanics methods using the MMFF force field of the MacroModel (V6.5) software package on the SGI O2 workstation [56].
Molecules 26 05292 g002
Scheme 2. Synthesis of the calix[4]arene-capped porphyrin (8).
Scheme 2. Synthesis of the calix[4]arene-capped porphyrin (8).
Molecules 26 05292 sch002
Scheme 3. Ion-binding ability of the calix[4]arene-capped porphyrin (11).
Scheme 3. Ion-binding ability of the calix[4]arene-capped porphyrin (11).
Molecules 26 05292 sch003
Scheme 4. Synthesis of the calix[5]arene-capped Zn-porphyrin 12.
Scheme 4. Synthesis of the calix[5]arene-capped Zn-porphyrin 12.
Molecules 26 05292 sch004
Scheme 5. Synthesis of the calix[4]arene-porphyrin 15.
Scheme 5. Synthesis of the calix[4]arene-porphyrin 15.
Molecules 26 05292 sch005
Scheme 6. (a) NBS, acetone, (b) K2CO3, PrI, acetone, (c) t-BuLi, THF, −78 °C.
Scheme 6. (a) NBS, acetone, (b) K2CO3, PrI, acetone, (c) t-BuLi, THF, −78 °C.
Molecules 26 05292 sch006
Scheme 7. Synthesis of the tiacalix[4]arene-porphyrin 17.
Scheme 7. Synthesis of the tiacalix[4]arene-porphyrin 17.
Molecules 26 05292 sch007
Figure 3. 1H-NMR titration curve of receptor 17 with the C70 fullerene in toluene (298 K, 300 MHz) (a) and structure of the 1:1 17-C70 complex (b) [36].
Figure 3. 1H-NMR titration curve of receptor 17 with the C70 fullerene in toluene (298 K, 300 MHz) (a) and structure of the 1:1 17-C70 complex (b) [36].
Molecules 26 05292 g003
Figure 4. Structures of the Bis-porphyrins 2328.
Figure 4. Structures of the Bis-porphyrins 2328.
Molecules 26 05292 g004
Figure 5. Molecular modeling structure of 26-C60 showing the close approach of the 10,20-tert-butylphenyl substituents to C60 and the more distant approach of the 15-tert-butylphenyl substituent [38].
Figure 5. Molecular modeling structure of 26-C60 showing the close approach of the 10,20-tert-butylphenyl substituents to C60 and the more distant approach of the 15-tert-butylphenyl substituent [38].
Molecules 26 05292 g005
Scheme 8. Conditions of synthesis of porphyrin-calixarenes (3133): (i) Pd(PPh3)2Cl2, toluene/Net3 (1/1), 60 °C; 3 days; (ii) Pd(PPh3)4, ZnBr2, 2.4 equiv., THF, Net3 (200 eqiv), rt, 3 days; (iii) BrCH2CH2CH3 (10 eqiv), NaH (10 equiv), THF, reflux, 8 days.
Scheme 8. Conditions of synthesis of porphyrin-calixarenes (3133): (i) Pd(PPh3)2Cl2, toluene/Net3 (1/1), 60 °C; 3 days; (ii) Pd(PPh3)4, ZnBr2, 2.4 equiv., THF, Net3 (200 eqiv), rt, 3 days; (iii) BrCH2CH2CH3 (10 eqiv), NaH (10 equiv), THF, reflux, 8 days.
Molecules 26 05292 sch008
Scheme 9. Conditions of synthesis of the porphyrin-calixarenes 37, 39 and 42: (i) Cs2Co3, CH3CN, reflux; (ii) 10% Pd(PPh3)2Cl2, CuI, Net3, 323K; (iii) NaI, DMF, 298 K.
Scheme 9. Conditions of synthesis of the porphyrin-calixarenes 37, 39 and 42: (i) Cs2Co3, CH3CN, reflux; (ii) 10% Pd(PPh3)2Cl2, CuI, Net3, 323K; (iii) NaI, DMF, 298 K.
Molecules 26 05292 sch009
Figure 6. Structures of the Bis-porpjhyrins 4345.
Figure 6. Structures of the Bis-porpjhyrins 4345.
Molecules 26 05292 g006
Scheme 10. Synthesis of the calix[4]arene-porphyrin conjugates 44 and 47.
Scheme 10. Synthesis of the calix[4]arene-porphyrin conjugates 44 and 47.
Molecules 26 05292 sch010
Scheme 11. Synthesis of the Zn-Bis-porphyrin-calix[4]arene conjugate 45.
Scheme 11. Synthesis of the Zn-Bis-porphyrin-calix[4]arene conjugate 45.
Molecules 26 05292 sch011
Scheme 12. (a) Synthesis of the Bis-porphyrin-calix[4]arene-crawn-[6] (49); (b) Synthesis of the ethoxycarbonyl-substituted Bis-porphyrin-calix[4]arene (50).
Scheme 12. (a) Synthesis of the Bis-porphyrin-calix[4]arene-crawn-[6] (49); (b) Synthesis of the ethoxycarbonyl-substituted Bis-porphyrin-calix[4]arene (50).
Molecules 26 05292 sch012
Scheme 13. 1:2 interaction of the receptor 50 with substrate DABCO.
Scheme 13. 1:2 interaction of the receptor 50 with substrate DABCO.
Molecules 26 05292 sch013
Figure 7. 1H-NMR spectra of external (51) (a) and internal (53), (b) complexes in CDCl3 (300 MHz, 298 K).
Figure 7. 1H-NMR spectra of external (51) (a) and internal (53), (b) complexes in CDCl3 (300 MHz, 298 K).
Molecules 26 05292 g007
Scheme 14. 1:1 interaction between the receptor 52 and DABCO in the presence of Na+.
Scheme 14. 1:1 interaction between the receptor 52 and DABCO in the presence of Na+.
Molecules 26 05292 sch014
Scheme 15. Reagents and conditions: (i) aminoporphyrin, Et3N/THF, rt; (ii) DCC/aminoporphyrin/CH2Cl2, rt.
Scheme 15. Reagents and conditions: (i) aminoporphyrin, Et3N/THF, rt; (ii) DCC/aminoporphyrin/CH2Cl2, rt.
Molecules 26 05292 sch015
Scheme 16. Reagents and conditions of the reaction of Bis-porphyrin-calixarene synthesis: (i) Excess pyrrole, rt/3 h; (ii) CHCl3, cat. BF3·Et2O, rt/7 h, DDQ, rt.; (iii) CH3CH2COOH, reflux.
Scheme 16. Reagents and conditions of the reaction of Bis-porphyrin-calixarene synthesis: (i) Excess pyrrole, rt/3 h; (ii) CHCl3, cat. BF3·Et2O, rt/7 h, DDQ, rt.; (iii) CH3CH2COOH, reflux.
Molecules 26 05292 sch016
Scheme 17. Synthesis of the calix[4]pyrrole-Bis-porphyrin 63.
Scheme 17. Synthesis of the calix[4]pyrrole-Bis-porphyrin 63.
Molecules 26 05292 sch017
Scheme 18. 1:1 complexation of the receptor 63 with F.
Scheme 18. 1:1 complexation of the receptor 63 with F.
Molecules 26 05292 sch018
Figure 8. Changes in the absorption spectra (a) and the titration curve (b) of receptor 63 with a fluoride ion [28] = 4.5 × 10−6 M, [F] = 0 ÷ 1,2 × 10−5 M in CD2Cl2, 298 K [67].
Figure 8. Changes in the absorption spectra (a) and the titration curve (b) of receptor 63 with a fluoride ion [28] = 4.5 × 10−6 M, [F] = 0 ÷ 1,2 × 10−5 M in CD2Cl2, 298 K [67].
Molecules 26 05292 g008
Figure 9. 1H-NMR spectra of receptor 63 (a) and complex with fluorine 64, (b) in CDCl3 (300 MHz, 298 K) [67].
Figure 9. 1H-NMR spectra of receptor 63 (a) and complex with fluorine 64, (b) in CDCl3 (300 MHz, 298 K) [67].
Molecules 26 05292 g009
Scheme 19. 1:2 complexation of the receptor 63 with DABCO.
Scheme 19. 1:2 complexation of the receptor 63 with DABCO.
Molecules 26 05292 sch019
Scheme 20. 1:1 complexation of the receptor 63 with DABCO in the presence of F.
Scheme 20. 1:1 complexation of the receptor 63 with DABCO in the presence of F.
Molecules 26 05292 sch020
Figure 10. Structures of porphyrin salt, tetrathiafulvalenecalix[4]pyrrole (67), and Li+-C60.
Figure 10. Structures of porphyrin salt, tetrathiafulvalenecalix[4]pyrrole (67), and Li+-C60.
Molecules 26 05292 g010
Figure 11. Supramolecular triad 68 composed of porphyrin carboxylate, radical cation 67, and Li+-C60 radical anion.
Figure 11. Supramolecular triad 68 composed of porphyrin carboxylate, radical cation 67, and Li+-C60 radical anion.
Molecules 26 05292 g011
Figure 12. (a) Nanosecond flash photolysis absorption spectra for a mixture of porphyrin carboxylate (100 mM), 67 (100 mM), and Li+-C60 (100 mM) following excitation at 532 nm in deoxygenated PhCN at 298 K as recorded 1 ms (black), 5 ms (red), and 20 ms (blue) after laser pulse irradiation. (b) Decay time profiles of the absorbance at 650 nm obtained using various laser intensities [9].
Figure 12. (a) Nanosecond flash photolysis absorption spectra for a mixture of porphyrin carboxylate (100 mM), 67 (100 mM), and Li+-C60 (100 mM) following excitation at 532 nm in deoxygenated PhCN at 298 K as recorded 1 ms (black), 5 ms (red), and 20 ms (blue) after laser pulse irradiation. (b) Decay time profiles of the absorbance at 650 nm obtained using various laser intensities [9].
Molecules 26 05292 g012
Scheme 21. Synthesis of theoxacalix-[2]-arene-[2]-Bis-(Zn-porphyrin) 69.
Scheme 21. Synthesis of theoxacalix-[2]-arene-[2]-Bis-(Zn-porphyrin) 69.
Molecules 26 05292 sch021
Figure 13. The curve of 1H-NMR titration of receptor 69 by the C70 fullerene in C6D6 (300 MHz, 298 K) [44].
Figure 13. The curve of 1H-NMR titration of receptor 69 by the C70 fullerene in C6D6 (300 MHz, 298 K) [44].
Molecules 26 05292 g013
Scheme 22. Synthesis of the resorcinarene-porphyrin supramolecular capsule (72).
Scheme 22. Synthesis of the resorcinarene-porphyrin supramolecular capsule (72).
Molecules 26 05292 sch022
Figure 14. (A) 1H-NMR spectrum of porphyrin 73 (a), resorcinarene 74 (b), capsule 72 (c), and its complex(with ethylene (d) at 298 K; (B) difference in the chemical shifts, Δδ Hi (·), and binding constants, K (O), depending on the substrate size. Δδ Hi = (Hi for the incapsulated substrate)-(Hi for the unbound substrate) [79].
Figure 14. (A) 1H-NMR spectrum of porphyrin 73 (a), resorcinarene 74 (b), capsule 72 (c), and its complex(with ethylene (d) at 298 K; (B) difference in the chemical shifts, Δδ Hi (·), and binding constants, K (O), depending on the substrate size. Δδ Hi = (Hi for the incapsulated substrate)-(Hi for the unbound substrate) [79].
Molecules 26 05292 g014
Scheme 23. Synthesis of the alkoxy substituted porphyrins 7577.
Scheme 23. Synthesis of the alkoxy substituted porphyrins 7577.
Molecules 26 05292 sch023
Figure 15. Structures of the ligands (Im, PrA, DAPr, DABCO).
Figure 15. Structures of the ligands (Im, PrA, DAPr, DABCO).
Molecules 26 05292 g015
Figure 16. Complexes of porphyrinate 75 and atropoisomers (a, b) of porphyrinate 76 with PrA.
Figure 16. Complexes of porphyrinate 75 and atropoisomers (a, b) of porphyrinate 76 with PrA.
Molecules 26 05292 g016
Figure 17. Complexes of dialkoxy-substituted porphyrinates 76 and 77 with DAPr [104].
Figure 17. Complexes of dialkoxy-substituted porphyrinates 76 and 77 with DAPr [104].
Molecules 26 05292 g017
Scheme 24. Synthesis of zinc porphyrinates with first-generation (79) and second-generation (82) carbazole branches.
Scheme 24. Synthesis of zinc porphyrinates with first-generation (79) and second-generation (82) carbazole branches.
Molecules 26 05292 sch024
Scheme 25. Synthesis of zinc porphyrinates with first-generation (79) and second-generation (82) carbazole branches.
Scheme 25. Synthesis of zinc porphyrinates with first-generation (79) and second-generation (82) carbazole branches.
Molecules 26 05292 sch025
Scheme 26. Complexation of the receptor 79 with TrA.
Scheme 26. Complexation of the receptor 79 with TrA.
Molecules 26 05292 sch026
Figure 18. Changes in the absorption spectra (a) and titration curves (b) of porphyrinate 79 and TrA in toluene, 20 °C. [109] = 2.7 × 10−5 M, [TrA] = 0 ÷ 6 × 10−5 M.
Figure 18. Changes in the absorption spectra (a) and titration curves (b) of porphyrinate 79 and TrA in toluene, 20 °C. [109] = 2.7 × 10−5 M, [TrA] = 0 ÷ 6 × 10−5 M.
Molecules 26 05292 g018
Figure 19. Structures of the meso-bis (84) and meso-tetra (85) carazolylphenyl substituted porphyrins.
Figure 19. Structures of the meso-bis (84) and meso-tetra (85) carazolylphenyl substituted porphyrins.
Molecules 26 05292 g019
Figure 20. Dependence of stability constants of porphyrinates 79, 8486 with Py, TrA, and DABCO on the porphyrinate and ligand nature.
Figure 20. Dependence of stability constants of porphyrinates 79, 8486 with Py, TrA, and DABCO on the porphyrinate and ligand nature.
Molecules 26 05292 g020
Scheme 27. Porphyrinates with two, four, and eight 4-carbazolylphenyl-1,2,3-triazole end groups.
Scheme 27. Porphyrinates with two, four, and eight 4-carbazolylphenyl-1,2,3-triazole end groups.
Molecules 26 05292 sch027
Figure 21. Changes in the absorption spectra (a) and titration curves (b) of porphyrinate 84 with DABCO in toluene, 20 °C. [84] = 1.0 × 10−6 M, [DABCO] = 0 ÷ 2.0 × 10−4 M.
Figure 21. Changes in the absorption spectra (a) and titration curves (b) of porphyrinate 84 with DABCO in toluene, 20 °C. [84] = 1.0 × 10−6 M, [DABCO] = 0 ÷ 2.0 × 10−4 M.
Molecules 26 05292 g021
Figure 22. Plot of the dependence of the optical density changes lg[(A0-Ai)/(Ai-Ak)] on the ligand concentration lgCL, C47 = 1.4 × 10−7 M, CDABCO = 3.0 × 10−7 ÷ 1.5 × 10−4 M.
Figure 22. Plot of the dependence of the optical density changes lg[(A0-Ai)/(Ai-Ak)] on the ligand concentration lgCL, C47 = 1.4 × 10−7 M, CDABCO = 3.0 × 10−7 ÷ 1.5 × 10−4 M.
Molecules 26 05292 g022
Figure 23. Conformational changes in cryptand-[2,2,2] during the formation of a complex with a potassium cation.
Figure 23. Conformational changes in cryptand-[2,2,2] during the formation of a complex with a potassium cation.
Molecules 26 05292 g023
Figure 24. Titration curves (a) and changes in the absorption spectra (b) of porphyrinate 82 during the formation of a complex with the Cr1 cryptand in toluene, 20°C. [82] = 1.5 × 10−6 M, [Cr1] = 0 ÷ 2.9 × 10−4 M.
Figure 24. Titration curves (a) and changes in the absorption spectra (b) of porphyrinate 82 during the formation of a complex with the Cr1 cryptand in toluene, 20°C. [82] = 1.5 × 10−6 M, [Cr1] = 0 ÷ 2.9 × 10−4 M.
Molecules 26 05292 g024
Scheme 28. Complexation of the receptor 79.
Scheme 28. Complexation of the receptor 79.
Molecules 26 05292 sch028
Figure 25. 1H-NMR spectrum of the 79-Cr1 complex in CDCl3 (300 MHz, 298 K); *—the signals of the equivalent (unsplit) central (-OCH2CH2O-, 12 H) protons, and nonequivalent (split) lateral [-OCH2-, (6H + 6H), and -CH2N-, (6H + 6H)] protons of the Cr1.
Figure 25. 1H-NMR spectrum of the 79-Cr1 complex in CDCl3 (300 MHz, 298 K); *—the signals of the equivalent (unsplit) central (-OCH2CH2O-, 12 H) protons, and nonequivalent (split) lateral [-OCH2-, (6H + 6H), and -CH2N-, (6H + 6H)] protons of the Cr1.
Molecules 26 05292 g025
Figure 26. Stability constants of the complexes of receptors 79, 82 and 86 with Cr1-Cr3 cryptands.
Figure 26. Stability constants of the complexes of receptors 79, 82 and 86 with Cr1-Cr3 cryptands.
Molecules 26 05292 g026
Scheme 29. Complexation of the receptor 79 with Cr1 in the presence of K+.
Scheme 29. Complexation of the receptor 79 with Cr1 in the presence of K+.
Molecules 26 05292 sch029
Figure 27. Changes in the absorption spectra in the [79-Cr1]–K+ system, [36] = 1.2 × 10−6 M, [L] = 0 ÷ 2.9 × 10−5 M (a); 1H-NMR spectrum of the [79-Cr1]–K+ system in CDCl3 (300 MHz, 298 K) (b); *—the signals of equivalent protons of the central (-OCH2CH2O-, 12 H) and lateral [-OCH2- (12 H), and -CH2N- (12 H)] groups.
Figure 27. Changes in the absorption spectra in the [79-Cr1]–K+ system, [36] = 1.2 × 10−6 M, [L] = 0 ÷ 2.9 × 10−5 M (a); 1H-NMR spectrum of the [79-Cr1]–K+ system in CDCl3 (300 MHz, 298 K) (b); *—the signals of equivalent protons of the central (-OCH2CH2O-, 12 H) and lateral [-OCH2- (12 H), and -CH2N- (12 H)] groups.
Molecules 26 05292 g027
Figure 28. Destruction of the 79-Cr1 (a) and 79-Cr2 (b) complex following the interaction with LiBr, NaBr, and KBr in a toluene–methanol (9:1) mixture.
Figure 28. Destruction of the 79-Cr1 (a) and 79-Cr2 (b) complex following the interaction with LiBr, NaBr, and KBr in a toluene–methanol (9:1) mixture.
Molecules 26 05292 g028
Scheme 30. Destruction of the 79-Cr1 complex in the presence of H+.
Scheme 30. Destruction of the 79-Cr1 complex in the presence of H+.
Molecules 26 05292 sch030
Scheme 31. Synthesis of the two-layered dendrimers 87G1 and 87G2.
Scheme 31. Synthesis of the two-layered dendrimers 87G1 and 87G2.
Molecules 26 05292 sch031
Figure 29. The scheme of selective binding of the fullerenes by 87G2 in absence (a) and presence (b) of the PhPy.
Figure 29. The scheme of selective binding of the fullerenes by 87G2 in absence (a) and presence (b) of the PhPy.
Molecules 26 05292 g029
Table 1. Bis-porphyrin “molecular tweezers” for binding of substrates of different natures [36,37,38,39,40,41,42,43,44,45,46,67,68,69,70,71,72,73,74,75,76].
Table 1. Bis-porphyrin “molecular tweezers” for binding of substrates of different natures [36,37,38,39,40,41,42,43,44,45,46,67,68,69,70,71,72,73,74,75,76].
“Molecular Tweezers”SubstratesRef.“Molecular Tweezers”SubstratesRef.
MPorph.LinkedMPorph.Linked
H2Porph.1
 
 
ZnPorph.2
Molecules 26 05292 i001 Molecules 26 05292 i002[36]
 
 
[37]
H2Porph.1 Molecules 26 05292 i003 Molecules 26 05292 i004[38]
H2Porph.1 Molecules 26 05292 i005 Molecules 26 05292 i006[38]H2Porph.1
H2Porph.3
H2Porph.4
Molecules 26 05292 i007 Molecules 26 05292 i008[38]
ZnPorph.6 Molecules 26 05292 i009 Molecules 26 05292 i010[39]ZnPorph.7 Molecules 26 05292 i011 Molecules 26 05292 i012[40]
ZnPorph.7 Molecules 26 05292 i013 Molecules 26 05292 i014[41]H2Porph.8 Molecules 26 05292 i015Regulation of the interporphyrin distance by the introduction of metal cations[42]
ZnPorph.3 Molecules 26 05292 i016 Molecules 26 05292 i017[43]ZnPorph.3 Molecules 26 05292 i018 Molecules 26 05292 i019[44]
H2Porph.9 Molecules 26 05292 i020Anions
Cl, Br, I, NO3
[45]H2Porph.10
ZnPorph.10
Molecules 26 05292 i021Fluorescent detection of Cu(II), Al(III), Cr(III), Zn(II), Cd(II)[46]
ZnPorph.11 Molecules 26 05292 i022 Molecules 26 05292 i023[67]ZnPorph.9 Molecules 26 05292 i024 Molecules 26 05292 i025[68]
ZnPorph.3 Molecules 26 05292 i026H2Porph.12[71]ZnPorph.3 Molecules 26 05292 i027H2Porph.12[70]
ZnPorph.6 Molecules 26 05292 i028DABCO[71]ZnPorph.9 Molecules 26 05292 i029DABCO[72]
ZnPorph.13 Molecules 26 05292 i030 Molecules 26 05292 i031[73]ZnPorph.14 Molecules 26 05292 i032 Molecules 26 05292 i033[74]
Bis-porphyrin “molecular tweezers” for recognition of chiral guest molecules.
ZnPorph.9 Molecules 26 05292 i034 Molecules 26 05292 i035[75]ZnPorph.9 Molecules 26 05292 i036 Molecules 26 05292 i037[76]
ZnPorph.15
MgPorph.15
Molecules 26 05292 i038 Molecules 26 05292 i039[15]ZnPorph.15
MgPorph.15
Molecules 26 05292 i040 Molecules 26 05292 i041[15]
ZnPorph.14 Molecules 26 05292 i042 Molecules 26 05292 i043[15]ZnPorph.9 Molecules 26 05292 i044 Molecules 26 05292 i045[15]
Table 2. Structure of the porphyrin fragments.
Table 2. Structure of the porphyrin fragments.
MPorph.R5, R10,R15R20R2,3,7,8,12,13,17,18MPorph.R5, R10,R15R20R2,3,7,8,12,13,17,18 Molecules 26 05292 i046
14-Me-PhPhH
24-Me-Ph9PhH
33,5-But-PhPh10H2,3,7,13,17,18-H, 8,12-n-C5H14
4(2,3,4,5,6-F-Ph)11Et
52,4,6-Me-Ph134-COOH-PhPhH
65,15-H;10-(3,5But-Ph)2,8,12,18-Me, 3,7,13,17-Bu125,10-Py, 15-Ph
72,15-H, 10-Ph2,8,12,18-Me,3,7,13,17-Et14Ph
8H2,3,17,18-H, 7,8,12,13-Et15H2,8,12,18-Me, 3,7,13,17-Et
Table 3. Stability constants, Ka (M−1), and energies of porphyrinate 7 binding of the substrate molecules in chloroform at 298 K, ΔG (kJ mol−1) [56].
Table 3. Stability constants, Ka (M−1), and energies of porphyrinate 7 binding of the substrate molecules in chloroform at 298 K, ΔG (kJ mol−1) [56].
SubstrateKa−ΔG
Pyridine (Py)53,400 ± 80027.0
4-methylpyridine (4-MePy)20 ± 17.4
Imidazole (Im)1,920,000 ± 40,00035.8
Table 4. Stability constants, Ka [M−1], ofthe 17-C60 and 17-C70 complexes at 25 °C [36].
Table 4. Stability constants, Ka [M−1], ofthe 17-C60 and 17-C70 complexes at 25 °C [36].
TolueneBenzene
13-M602.0 × 10313-M602.1 × 103
13-M703.6 × 10413-M702.7× 104
Table 5. Stability constants, Ka [M−1], of the 58-C70 and 60-C70 complexes (1H-NMR, 25 °C, 300 MHz) [37].
Table 5. Stability constants, Ka [M−1], of the 58-C70 and 60-C70 complexes (1H-NMR, 25 °C, 300 MHz) [37].
Toluene-d8Benzene-d6
58-M701300 ± 5001100 ± 200
60-M701000 ± 2004500 ± 600
Table 6. Stability constants, Ka [M−1], of 1:1 complexes of porphyrinates (79,8486) and organic ligands (Py, TrA) in toluene, CZnP ≈ 1.1 ÷ 1.6 × 10−5 M.
Table 6. Stability constants, Ka [M−1], of 1:1 complexes of porphyrinates (79,8486) and organic ligands (Py, TrA) in toluene, CZnP ≈ 1.1 ÷ 1.6 × 10−5 M.
PyTrA
864805800
84901200
852403900
79110,000970,000
Table 7. Constants of binding (lgKa) of alkali metal cations by cryptand-[2,2,2] in methanol at 25 °C [110].
Table 7. Constants of binding (lgKa) of alkali metal cations by cryptand-[2,2,2] in methanol at 25 °C [110].
Metal CationLi+Na+K+
Ionic radius, Ǻ1.361.902.66
Binding constant1.807.219.75
Table 8. Stability constants (lgKa) of the Cr1 complexes with 82 and 79 in tolueneanda (9:1) toluene–methanol mixture at 20 °C [110].
Table 8. Stability constants (lgKa) of the Cr1 complexes with 82 and 79 in tolueneanda (9:1) toluene–methanol mixture at 20 °C [110].
82-Cr1
(Toluene)
82-Cr1
(Toluene–Methanol)
79-Cr1
(Toluene)
79-Cr1
(Toluene–Methanol)
2.882.765.965.87
Table 9. Stability constants, Ka × 104 [M−1], of the 87G1 and 87G2 complexes with C60, C70, and C84 in a (2:1) toluene–acetonitrile mixture, 25 °C [112,113].
Table 9. Stability constants, Ka × 104 [M−1], of the 87G1 and 87G2 complexes with C60, C70, and C84 in a (2:1) toluene–acetonitrile mixture, 25 °C [112,113].
C60C70C84
87-G11.604.905.80
87-G22.106.005.40
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Mamardashvili, G.; Mamardashvili, N.; Koifman, O. Macrocyclic Receptors for Identification and Selective Binding of Substrates of Different Nature. Molecules 2021, 26, 5292. https://doi.org/10.3390/molecules26175292

AMA Style

Mamardashvili G, Mamardashvili N, Koifman O. Macrocyclic Receptors for Identification and Selective Binding of Substrates of Different Nature. Molecules. 2021; 26(17):5292. https://doi.org/10.3390/molecules26175292

Chicago/Turabian Style

Mamardashvili, Galina, Nugzar Mamardashvili, and Oscar Koifman. 2021. "Macrocyclic Receptors for Identification and Selective Binding of Substrates of Different Nature" Molecules 26, no. 17: 5292. https://doi.org/10.3390/molecules26175292

Article Metrics

Back to TopTop