Next Article in Journal
2,3,4,5,6-Pentabromobenzoic Acid
Previous Article in Journal
Methyl 5-Imino-2-methyl-1,10a-dihydro-5H-chromeno[2,3-b]pyridine-3-carboxylate
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Communication

Assembly of Sn(IV)-Porphyrin Cation Exhibiting Supramolecular Interactions of Anion···Anion and Anion···π Systems

Department of Chemistry and Bioscience, Kumoh National Institute of Technology, Gumi 39177, Korea
Molbank 2022, 2022(4), M1454; https://doi.org/10.3390/M1454
Submission received: 8 September 2022 / Revised: 22 September 2022 / Accepted: 22 September 2022 / Published: 25 September 2022

Abstract

:
Trans-diaqua[meso-tetrakis(4-pyridyl)porphyrinato]Sn(IV) dinitrate complexes were assembled in a two-dimensional manner via hydrogen bonding between aqua ligands and pyridyl substituents. Interestingly, this supramolecular assembly was accompanied by unconventional noncovalent interactions, such as anion···anion and anion···π interactions, which were confirmed by X-ray crystallographic analysis. Two nitrate anions close to 2.070 Å were constrained in a confined space surrounded by four hydrogen-bonded Sn(IV)-porphyrin cations. The nitrate anion was also 3.433 Å away from the adjacent pyrrole ring, and the dihedral angle between the two mean planes was estimated to be 7.39°. The preference of the anion···π interaction was related to the electron-deficient π-system owing to the high-valent Sn(IV) center and cationic nature of the porphyrin complex. These two unconventional noncovalent interactions played an important role in the formation of a one-dimensional array with pairs of Sn(IV)-porphyrin cation and nitrate anion.

Graphical Abstract

1. Introduction

The supramolecular assembly of metalloporphyrins provides valuable insights for elucidating biochemical processes and developing photoelectronic materials [1,2,3,4,5,6,7,8,9]. Noncovalent intermolecular interactions, such as metal–ligand coordination [10,11,12,13], hydrogen bonding [14,15], and π-stacking [16,17,18], have been mainly utilized in the supramolecular assembly of metalloporphyrins. Introduction of specific functional groups onto the peripheral positions enables the corresponding intermolecular interactions. In this context, Sn(IV)-porphyrin is a more attractive motif for constructing self-assembled metalloporphyrin nanostructures [19,20,21,22], because they can exhibit unique features by strongly adopting functional ligands at their axial positions [23,24]. Thus, cooperative interactions between two different functional groups in six-coordinated Sn(IV)-porphyrins have led to various assembled structures for the development of functional porphyrin materials [25,26,27,28,29,30,31,32,33]. Recently, nanomaterials based on the assembly of Sn(IV)-porphyrins have been intensively studied for application to visible light-sensitive photocatalysis [34,35,36,37,38,39,40,41].
Among the various types of supramolecular assemblies, ionic assemblies are accompanied by the coupling of structurally different ionic building blocks via strong electrostatic interactions [42]. The predominant driving force arises from the attractive electrostatic interaction between cation and anion species. Accordingly, anions and cations are arranged in a way that minimizes the distance between oppositely charged species, whereas electrostatic repulsion maximizes the space between like-charged species (cation···cation or anion···anion). However, certain ionic assembled structures involve unconventional interactions, indicated by the unusual proximity between like-charged species [43,44,45]. Moreover, these types of contacts can play an important role in the formation of specific supramolecular assemblies [46,47,48,49].
In contrast, porphyrin macrocycles are composed of well-defined and largely delocalized π-electron systems. The π-electronic nature of porphyrin molecules crucially contributes to the supramolecular assembly through π-stacking [16,17,18]. Owing to their electron-deficient character, cation species can attractively interact with π-electron systems, commonly known as cation···π interactions [50,51,52]. However, anion···π interactions are unconventional because anions and π-electron systems are expected to interact repulsively with each other due to the involvement of electron-rich entities [53,54]. In stark contrast to the mature research field regarding cation···π interaction, anion···π interaction remains a challenging area to explore in supramolecular chemistry [55].
The supramolecular assembly of an amphoteric Sn(IV)-porphyrin cation complex is reported here, in which unconventional noncovalent interactions such as anion···anion and anion···π notably occurred and played an important role in constructing a specific supramolecular array. Sn(IV)-porphyrin cation complexes exhibiting unconventional interactions of anion···anion and anion···π could provide new insights for creating elaborate ionic assemblies.

2. Results and Discussion

Trans-diaqua[meso-tetrakis(4-pyridyl)porphyrinato]Sn(IV) dication complex (SnP2+) was prepared using an acid–base equilibrium reaction between trans-dihydroxo[meso-tetrakis(4-pyridyl)porphyrinato]Sn(IV) (SnP) [56] and trans-diaqua[meso-tetrakis(4-pyridinium)porphyrinato]Sn(IV) hexacation complex (SnP6+) [57] (Scheme 1). SnP has six Lewis-base atoms from the hydroxo ligand and pyridyl substituent, whereas SnP6+ has six acidic hydrogen atoms. Accordingly, the two complementary complexes produced an amphoteric complex, SnP2+, via the acid–base reaction.
In the ESI mass spectrum, a parent ion peak appeared at m/z 753.1, which is consistent with the formula of the [SnP2+–OH]+ species (Figure S1a). A molecular ion peak corresponding to NO3 was also observed in the negative mode of the ESI mass spectrum (Figure S1b). In the 1H NMR spectrum of SnP2+ in DMSO-d6 (Figure S2), the resonances for the pyridyl and β-pyrrolic protons were notably shifted to a lower field compared to SnPδ = 0.07 (Hpy-α), 0.08 (Hpy-β), and 0.30 (Hpyrrole) ppm). However, compared to the six-valent charged analog, SnP6+, the resonances for SnP2+ were oppositely shifted to a higher field (Δδ = −0.06 (Hpy-α), −0.11 (Hpy-β), and −0.05 (Hpyrrole) ppm). Therefore, the 1H NMR spectral feature of SnP2+ indicated that the electronic characteristics of SnP2+ were intermediate between SnP and SnP6+. In the comparison of FT-IR spectra of SnP and SnP2+ (Figure S3), the characteristic band for the NO3 anion (1350 cm−1) was found only in SnP2+.
The hydrogen-bonded supramolecular assembly of SnP [56] and the supramolecular assembly of SnP6+ stabilized by ionic hydrogen-bonding interactions [57] were previously reported, and have been intensively studied, including X-ray crystallography. SnP or SnP6+ provides a unilateral component for hydrogen bonding. SnP and SnP6+ have hydrogen-bonding acceptors (hydroxo ligand and pyridyl substituent) and donors (aqua ligand and pyridinium substituent), respectively. Furthermore, SnP2+ can perform both functions (aqua ligand as the donor and a pyridyl substituent as the acceptor). Therefore, the supramolecular assembly of SnP2+ based on hydrogen bonding and ionic interactions is of interest. A single crystal suitable for X-ray crystallographic analysis was obtained from the bulk as-synthesized samples. SnP2+ crystallized into a monoclinic system with a P21/c space group. Crystallographic analysis identified three chemical species: (i) trans-diaqua[meso-tetrakis(4-pyridyl)porphyrinato]Sn(IV) cation, (ii) nitrate anion, and (iii) clathrate acetone molecules, which were formulated as [Sn(OH2)2(TPyP)](NO3)2·2acetone.
Figure 1 shows the structure of SnP2+ cation in the crystal. The central Sn(IV) ion adopts an octahedral geometry with the porphyrin occupying the square base and the axial coordination of two aqua ligands. The Sn(IV) ion lies on an inversion center, so that the two aqua ligands lie in an exact trans orientation with respect to each other. The average Sn–N bond distance was estimated to be 2.091(4) Å, which was shorter than that of SnP (2.101(3) Å) and longer than that of SnP6+ (2.083(2) Å). The aqua ligand in SnP2+ was clearly distinguished from the hydroxo ligand coordinated to Sn(IV)-porphyrin by comparing their bond distances. The Sn–Oaqua bond distance of 2.093(2) Å in SnP2+ is significantly longer than the Sn–Ohydroxo bond distance of 2.014(3) Å in SnP, but is similar to that of SnP6+ (2.1024(19) Å). The structural parameters related to the Sn(IV) center in SnP2+ are summarized in Table 1 and compared with those for SnP6+ and SnP.
The aqua ligand of SnP2+ was hydrogen-bonded to the pyridyl nitrogen atom of an adjacent Sn(IV)-porphyrin cation, as shown in Figure 2a. The structural parameters associated with hydrogen bonding are summarized in Table 2. The distance from the aqua oxygen atom to the hydrogen-bonded pyridyl nitrogen atom was estimated to be 2.695(4) Å (O1W···N3) including 1.90 Å of H1WA···N3. The three atoms for hydrogen bonding (O1W–H1WA···N3) were arranged almost linearly at an angle of 165.2°. Thus, the Sn(IV)-porphyrin building blocks were assembled in a two-dimensional (2D) manner through hydrogen bonding between the aqua ligand and the pyridyl substituent. Figure 2b shows a hydrogen-bonded array of Sn(IV)-porphyrin cations.
In contrast, two nitrate anions reside in a confined space surrounded by four Sn(IV)-porphyrin cations, as shown in Figure 3a. The nitrate anion, which acts as a hydrogen-bonding acceptor, is also structurally linked with the Sn(IV)-porphyrin cation. The aqua ligand hooked the nitrate anion via hydrogen bonding in the supramolecular assembly of SnP2+ (Figure 3b). The distance from the nitrate oxygen atom to the hydrogen-bonded aqua oxygen atom is estimated to be 2.610(7) Å (O1W···O2) including 1.81 Å of H1WB···O2. The three atoms for hydrogen bonding (O1W–H1WB···O2) were arranged almost linearly at an angle of 167.6°.
More importantly, the two adjacent nitrate anions are unusually close to each other despite the repulsive forces between the same charges. The nitrate anion was 2.070 Å (O3···O3’) apart from the other nitrate anions, as shown in Figure 4a. The additional structural parameters for this contact were 2.435 Å for O3···O4’ and 2.688 Å for O3···N5’. The two nitrate anions were also coplanar with a dihedral angle of 0°. A similar unusual proximity around nitrate anions was also found in the literature [58,59,60], including 2.101(6) Å of the O···O distance between two nitrate anions in a lattice of {[Cd(BMIB)(H2O)2](NO3)2} (BMID = 1,4-bis(4-methyl-imidazolyl)benzene) [58]. The close contact between these two nitrate anions means that significant anion···anion interactions have occurred. Forced contact in a confined space and stabilization by the hydrogen bond with the aqua ligand resulted in an unconventional anion···anion interaction. Unusual short distances of like-charged ions may occur because of either packing constraints or additional attractive interactions, which stabilize the crystal structure [43,44,45].
Another peculiarity is the occurrence of a nitrate···π-system of pyrrole contacts in the crystal of SnP2+ (Figure 4b,c). The centroid of the nitrate anion was 3.433 Å from that of an adjacent pyrrole ring, and the dihedral angle between the two mean planes was estimated to be 7.39°. The high-valent Sn(IV) center and the additional cationic nature of SnP2+ made the porphyrin ligand an electron-deficient π-system. Consequently, the nitrate anion preferentially interacted with the pyrrole ring. Notably, this is a rare example of crystallographic evidence of the simultaneous occurrence of anion···anion and anion···π interactions.
The Sn(IV)-porphyrin cation and nitrate anion layers were alternately stacked to form a one-dimensional (1D) supramolecular array (Figure 5a). Given the formally repulsive nature of contact between two nitrate anions, this assembly was expected to have extremely low energy compared to a cation···anion array. However, crystal packing was partially affected by this weak interaction. Consequently, unconventional anion···anion interactions between the two nitrate anions played a critical role in forming the 1D array with pairs of Sn(IV)-porphyrin cation and nitrate anion. There were two different directions for the 1D arrays, and these 1D arrays expanded alternately, as shown in Figure 5b. In adjacent stacked 1D arrays with different orientations, Sn(IV)-porphyrins were tilted at a dihedral angle of approximately 76.6° with respect to each other.

3. Materials and Methods

Trans-dihydroxo[meso-tetrakis(4-pyridyl)porphyrinato]Sn(IV) (SnP) [56] and trans-diaqua[meso-tetrakis(4-pyridinium)porphyrinato]Sn(IV) (SnP6+) hexanitrate [57] complexes were prepared using previously reported procedures. UV-vis and IR spectra were recorded using a Hewlett-Packard 8453 diode array and a Jasco FTIR-460 Plus spectrophotometer, respectively. 1H NMR spectra were obtained using a Bruker BIOSPIN/AVANCE III 400 spectrometer. Electrospray ionization (ESI) mass spectra were recorded using a Waters ZQ 2000 LC/MS spectrometer. Elemental analyses were performed using a ThermoQuest EA 1110 analyzer.

3.1. Preparation of Trans-diaqua[meso-tetrakis(4-pyridyl)porphyrinato]Sn(IV) Dinitrate

Trans-dihydroxo[meso-tetrakis(4-pyridyl)porphyrinato]Sn(IV) (385.0 mg, 0.5 mmol) and trans-diaqua[meso-tetrakis(4-pyridinium)porphyrinato]Sn(IV) hexanitrate (287.0 mg, 0.25 mmol) were mixed in 20 mL of water. The reaction mixture was heated under reflux for 24 h and stirred at room temperature for an additional 24 h. After the acetone (100 mL) was layered over for slow diffusion, the solution was allowed to stand for a week. Block-shaped violet crystals were obtained via filtration. Yield: 86%. 1H NMR (400 MHz, DMSO-d6) δ 9.42 (s, 8H, Hpyrrole), 9.18 (d, J = 2.2 Hz, 8H, Hpy-α), 8.38 (d, J = 4.1 Hz, 8H, Hpy-β). UV-vis (DMSO, nm): λmax (log ε) 410 (4.20), 520 (3.05), 552 (3.78), and 592 (3.52). Mass (ESI): m/z 754.0 [(M–H2O)+ requires 754.13]. Elemental analysis was performed on evacuated samples. Anal. Calcd. for C40H28N10O8Sn: C, 53.65; H, 3.15; N, 15.64. Found: C, 53.23; H, 3.58; and N, 15.21.

3.2. X-ray Crystal Structure Determination

Single crystals in the as-synthesized sample were picked up using Paraton-N hydrocarbon oil on a glass fiber and mounted on a Bruker APEX-II CCD diffractometer equipped with a graphite monochromated Mo Kα (λ = 0.71073 Å) radiation source and a CCD detector. Data were collected at 173(2) K in a cold N2 stream. Three sets of 15 frames of two-dimensional diffraction images were collected and processed to deduce the cell parameter and orientation matrix. All crystallographic data were corrected for Lorentz and polarization effects (SAINT) [61]. Empirical absorption correction was applied (SADABS) [62]. The integration of data using triclinic unit cell yielded 12,463 reflections, of which 4,495 were independent and greater than 2σ(I). The final cell constants were refined with 4,495 reflections at 2.059° ≤ θ ≤ 26.998°. Data analysis showed negligible decay during data collection. The structure was solved using direct methods and the subsequent difference Fourier syntheses and refined using the SHELXTL software package [63]. Crystallographic data and additional details of data collection and refinement are summarized in Table S1. All non-hydrogen atoms were refined anisotropically. All hydrogen atoms were included in the calculated positions, with isotropic thermal parameters 1.2 times those of the attached atoms. The final full matrix least-squares refinement of F2 converged to R1 = 0.0471 (all data) and wR2 = 0.1192 (I > 2σ(I)) with GOF = 1.116.

4. Conclusions

SnP2+ was synthesized utilizing the acid–base reaction of Lewis bases in SnP and protons in SnP6+, demonstrating that two complementary porphyrin complexes can generate an amphoteric porphyrin complex from the acid–base reaction. SnP2+ has components for hydrogen bonding, with aqua ligands and pyridyl substituents serving as the donor and acceptor, respectively. X-ray structural analysis revealed that the Sn(IV)-porphyrin building blocks were assembled in a 2D manner through hydrogen bonding between the aqua ligand and pyridyl substituent. Furthermore, the aqua ligand simultaneously interacted with the nitrate anion via hydrogen bonding in the supramolecular assembly of SnP2+. More importantly, the unconventional anion···anion interactions of the two nitrate anions occurred in a confined space surrounded by four Sn(IV)-porphyrin cations. The anion···π-system interaction, another unconventional interaction, appeared in the proximity of the nitrate anion with the pyrrole ring. This is a rare example in which anion···anion and anion···π interactions occurred simultaneously. These findings can provide new opportunities in supramolecular chemistry, particularly for molecular hosts composed of oligomeric Sn(IV)-porphyrin with multivalent cationic charges to recognize polyanions in confined spaces.

Supplementary Materials

The following are available online: Figure S1. ESI mass spectra of SnP2+ (a) positive mode, (b) negative mode; Figure S2. 1H NMR spectrum of SnP2+ in DMSO-d6; Figure S3. Comparison of FT-IR spectra of SnP and SnP2+ (KBr pellet); Table S1. Crystal and structure refinement data.

Funding

This work was supported by the Kumoh National Institute of Technology (2021).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

CCDC 2173158 contains supplementary crystallographic data for this paper. These data can be obtained free of charge via http://www.ccdc.cam.ac.uk/conts/retrieving.html (or from the CCDC, 12 Union Road, Cambridge, CB2 1EZ, UK; Fax: +44 1223 336033; E-mail: [email protected]).

Acknowledgments

The author thanks Jaheon Kim for his helpful discussion.

Conflicts of Interest

The author declares no conflict of interest.

Sample Availability

Samples of the compounds are not available from the authors.

References

  1. Drain, C.M.; Hupp, J.T.; Suslick, K.S.; Wasielewski, M.R.; Chen, X. A perspective on four new porphyrin-based functional materials and devices. J. Porphyr. Phthalocyanines 2002, 6, 243–258. [Google Scholar] [CrossRef]
  2. Milic, T.N.; Chi, N.; Yablon, D.G.; Flynn, G.W.; Batteas, J.D.; Drain, C.M. Controlled hierarchical self-assembly and deposition of nanoscale photonic materials. Angew. Chem. Int. Ed. 2002, 41, 2117–2119. [Google Scholar] [CrossRef]
  3. Suslick, K.S.; Rakow, N.A.; Kosal, M.E.; Chou, J.-H. The materials chemistry of porphyrins and metalloporphyrins. J. Porphyr. Phthalocyanines 2000, 4, 407–413. [Google Scholar] [CrossRef]
  4. Chambron, J.-C.; Heitz, V.; Sauvage, J.-P. The Porphyrin Handbook; Kadish, K.M., Smith, K.M., Guilard, R., Eds.; Academic Press: San Diego, CA, USA, 2000; Volume 6, pp. 1–42. [Google Scholar]
  5. Chou, J.-H.; Nalwa, H.S.; Kosal, M.E.; Rakow, N.A.; Suslick, K.S. Applications of Porphyrins and Metalloporphyrins to Materials Chemistry. In The Porphyrin Handbook; Kadish, K.M., Smith, K.M., Guilard, R., Eds.; Academic Press: San Diego, CA, 2000; Volume 6, pp. 43–131. [Google Scholar]
  6. George, S.; Goldberg, I. Self-Assembly of Supramolecular Porphyrin Arrays by Hydrogen Bonding: New Structures and Reflections. Cryst. Growth Des. 2006, 6, 755–762. [Google Scholar] [CrossRef]
  7. Vinodu, M.; Goldberg, I. New assembly modes of porphyrin-based networks. CrystEngComm 2003, 5, 204–207. [Google Scholar] [CrossRef]
  8. Goldberg, I. Crystal engineering of porphyrin framework solids. Chem. Commun. 2005, 1243–1254. [Google Scholar] [CrossRef]
  9. Hasobe, T. Photo-and electro-functional self-assembled architectures of porphyrins. Phys. Chem. Chem. Phys. 2012, 14, 15975–15987. [Google Scholar] [CrossRef]
  10. Northrop, B.H.; Zheng, Y.-R.; Chi, K.-W.; Stang, P.J. Self-Organization in Coordination-Driven Self-Assembly. Acc. Chem. Res. 2009, 42, 1554–1563. [Google Scholar] [CrossRef]
  11. Horike, S.; Shimomura, S.; Kitagawa, S. Soft porous crystals. Nat. Chem. 2009, 1, 695–704. [Google Scholar] [CrossRef]
  12. Farha, O.K.; Shultz, A.M.; Sarjeant, A.A.; Nguyen, S.T.; Hupp, J.T. Active-Site-Accessible, Porphyrinic Metal-Organic Framework Materials. J. Am. Chem. Soc. 2011, 133, 5652–5655. [Google Scholar] [CrossRef]
  13. Lee, C.Y.; Farha, O.K.; Hong, B.J.; Sarjeant, A.A.; Nguyen, S.T.; Hupp, J.T. Light-Harvesting Metal-Organic Frameworks (MOFs): Efficient Strut-to-Strut Energy Transfer in Bodipy and Porphyrin-Based MOFs. J. Am. Chem. Soc. 2011, 133, 15858–15861. [Google Scholar] [CrossRef]
  14. Würthner, F.; Kaiser, T.E.; Saha-Möller, C.R. J-Aggregates: From Serendipitous Discovery to Supramolecular Engineering of Functional Dye Materials. Angew. Chem. Int. Ed. 2011, 50, 3376–3410. [Google Scholar] [CrossRef]
  15. Madueno, R.; Raisanen, M.T.; Silien, C.; Buck, M. Functionalizing hydrogen-bonded surface networks with self-assembled monolayers. Nature 2008, 454, 618–621. [Google Scholar] [CrossRef]
  16. Watson, M.D.; Fechtenkotter, A.; Müllen, K. Big Is Beautiful-“Aromaticity” Revisited from the Viewpoint of Macromolecular and Supramolecular Benzene Chemistry. Chem. Rev. 2001, 101, 1267–1300. [Google Scholar] [CrossRef]
  17. Hunter, C.A.; Sanders, J.K.M. The nature of .Pi.-.Pi. Interactions. J. Am. Chem. Soc. 1990, 112, 5525–5534. [Google Scholar] [CrossRef]
  18. Zang, L.; Che, Y.; Moore, J.S. One-Dimensional Self-Assembly of Planar π-Conjugated Molecules: Adaptable Building Blocks for Organic Nanodevices. Acc. Chem. Res. 2008, 41, 1596–1608. [Google Scholar] [CrossRef]
  19. Medforth, C.J.; Wang, Z.; Martin, K.E.; Song, Y.; Jacobsen, J.L.; Shelnutt, J.A. Self-assembled porphyrin nanostructures. Chem. Commun. 2009, 7261–7277. [Google Scholar] [CrossRef] [PubMed]
  20. Kim, H.-J.; Jo, H.J.; Kim, J.; Kim, S.-Y.; Kim, D.; Kim, K. Supramolecular self-assembly of tin(IV) porphyrin channels stabilizing single-file chains of water molecules. CrystEngComm 2005, 7, 417–420. [Google Scholar] [CrossRef]
  21. Shee, N.K.; Lee, C.-J.; Kim, H.-J. Hexacoordinated Sn(IV) porphyrin-based square-grid frameworks exhibiting selective uptake of CO2 over N2. Bull. Korean Chem. Soc. 2022, 43, 103–109. [Google Scholar] [CrossRef]
  22. Shee, N.K.; Jo, H.J.; Kim, H.-J. Coordination framework materials fabricated by the self-assembly of Sn(IV) porphyrins with Ag(I) ions for the photocatalytic degradation of organic dyes in wastewater. Inorg. Chem. Front. 2022, 9, 1270–1280. [Google Scholar] [CrossRef]
  23. Arnold, D.P.; Blok, J. The coordination chemistry of tin porphyrin complexes. Coord. Chem. Rev. 2004, 248, 299–319. [Google Scholar] [CrossRef]
  24. Sanders, J.K.M.; Bampos, N.; Clyde-Watson, Z.; Darling, S.L.; Hawley, J.C.; Kim, H.-J.; Mak, C.C.; Webb, S.J. The Porphyrin Handbook; Kadish, K.M., Smith, K.M., Guilard, R., Eds.; Academic Press: San Diego, CA, USA, 2000; Volume 3, pp. 1–48. [Google Scholar]
  25. Kim, H.-J.; Bampos, N.; Sanders, J.K.M. Assembly of Dynamic Heterometallic Oligoporphyrins Using Cooperative Zinc-Nitrogen, Ruthenium-Nitrogen, and Tin-Oxygen Coordination. J. Am. Chem. Soc. 1999, 121, 8120–8121. [Google Scholar] [CrossRef]
  26. Kim, H.J.; Park, K.-M.; Ahn, T.K.; Kim, S.K.; Kim, K.S.; Kim, D.; Kim, H.-J. Novel fullerene–porphyrin–fullerene triad linked by metal axial coordination: Synthesis, X-ray crystal structure, and spectroscopic characterizations of trans-bis([60]fullerenoacetato)tin(iv) porphyrin. Chem. Commun. 2004, 22, 2594–2595. [Google Scholar] [CrossRef] [PubMed]
  27. Kim, H.J.; Jeon, W.S.; Lim, J.H.; Hong, C.S.; Kim, H.-J. Synthesis, X-ray crystal structure, and electrochemistry of trans-bis(ferrocenecarboxylato)(tetraphenylporphyrinato)tin(IV). Polyhedron 2007, 26, 2517–2522. [Google Scholar] [CrossRef]
  28. Kim, H.J.; Jang, J.H.; Choi, H.; Lee, T.; Ko, J.; Yoon, M.; Kim, H.-J. Photoregulated Fluorescence Switching in Axially Coordinated Tin(IV) Porphyrinic Dithienylethene. Inorg. Chem. 2008, 47, 2411–2415. [Google Scholar] [CrossRef]
  29. Kim, S.H.; Kim, H.; Kim, K.; Kim, H.-J. The first tin(IV) porphyrin complex with chiral amino acid ligands: Synthesis, characterization and X-ray crystal structure of trans-bis(L-prolinato)[5,10,15,20-tetrakis-(4-pyridyl)porphyrinato]tin(IV). J. Porphyrins Phthalocyanines 2009, 13, 805–810. [Google Scholar] [CrossRef]
  30. Singh, A.P.; Park, B.B.; Kim, H.-J. Facile C-C bond cleavage of β-diketones by tin(IV) porphyrin complex. Tetrahedron Lett. 2012, 53, 6456–6459. [Google Scholar] [CrossRef]
  31. Kim, H.J.; Shee, N.K.; Park, K.-M.; Kim, H.-J. Assembly and X-ray crystal structures of heterometallic multiporphyrins with complementary coordination between ruthenium (II) and tin (IV) porphyrins. Inorg. Chim. Acta 2019, 488, 1–7. [Google Scholar] [CrossRef]
  32. Shee, N.K.; Kim, M.K.; Kim, H.-J. Fluorescent chemosensing for aromatic compounds by supramolecular complex composed of tin(IV) porphyrin, viologen, and cucurbit[8]uril. Chem. Commun. 2019, 55, 10575–10578. [Google Scholar] [CrossRef]
  33. Kim, M.K.; Shee, N.K.; Lee, J.; Yoon, M.; Kim, H.-J. Photoinduced Electron Transfer upon Supramolecular Complexation of (Porphyrinato) Sn-Viologen with Cucurbit [7] uril. Photochem. Photobiol. Sci. 2019, 18, 1996–2002. [Google Scholar] [CrossRef]
  34. Shee, N.K.; Kim, M.K.; Kim, H.-J. Supramolecular Porphyrin Nanostructures Based on Coordination-Driven Self-Assembly and Their Visible Light Catalytic Degradation of Methylene Blue Dye. Nanomaterials 2020, 10, 2314. [Google Scholar] [CrossRef] [PubMed]
  35. Shee, N.K.; Kim, H.-J. Self-Assembled Nanomaterials Based on Complementary Sn(IV) and Zn(II)-Porphyrins, and Their Photocatalytic Degradation for Rhodamine B Dye. Molecules 2021, 26, 3598. [Google Scholar] [CrossRef] [PubMed]
  36. Shee, N.K. Kim, H.-J. Three Isomeric Zn(II)-Sn(IV)-Zn(II) Porphyrin-Triad-Based Supramolecular Nanoarchitectures for the Morphology-Dependent Photocatalytic Degradation of Methyl Orange. ACS Omega 2022, 7, 9775–9784. [Google Scholar] [CrossRef] [PubMed]
  37. Jang, J.H.; Jeon, K.-S.; Oh, S.; Kim, H.-J.; Asahi, T.; Masuhara, H.; Yoon, M. Synthesis of Sn-porphyrin-intercalated trititanate nanofibers:  Optoelectronic properties and photocatalytic activities. Chem. Mater. 2007, 19, 1984–1991. [Google Scholar] [CrossRef]
  38. Kim, W.; Park, J.; Jo, H.J.; Kim, H.-J.; Choi, W. Visible light photocatalysts based on homogeneous and heterogenized tin porphyrins. J. Phys. Chem. C 2008, 112, 491–499. [Google Scholar] [CrossRef]
  39. Kim, W.; Tachikawa, T.; Majima, T.; Li, C.; Kim, H.-J.; Choi, W. Tin-porphyrin sensitized TiO2 for the production of H2 under visible light. Energy Environ. Sci. 2010, 3, 1789–1795. [Google Scholar] [CrossRef]
  40. Kim, H.; Kim, W.; Mackeyev, Y.; Lee, G.-S.; Kim, H.-J.; Tachikawa, T.; Hong, S.; Lee, S.; Kim, J.; Wilson, L.J.; et al. Selective oxidative degradation of organic pollutants by singlet oxygen-mediated photosensitization: Tin porphyrin versus C60 aminofullerene systems. Environ. Sci. Tech. 2012, 46, 9606–9613. [Google Scholar] [CrossRef]
  41. Li, C.; Park, K.-M.; Kim, H.-J. Ionic assembled hybrid nanoparticle consisting of tin(IV) porphyrin cations and polyoxomolybdate anions, and photocatalytic hydrogen production by its visible light sensitization. Inorg. Chem. Commun. 2015, 60, 8–11. [Google Scholar] [CrossRef]
  42. Faul, C.F.J.; Antonietti, M. Ionic self-assembly: Facile synthesis of supramolecular materials. Adv. Mater. 2003, 15, 673–683. [Google Scholar] [CrossRef]
  43. He, Q.; Tu, P.; Sessler, J.L. Supramolecular Chemistry of Anionic Dimers, Trimers, Tetramers and Clusters. Chem. 2018, 4, 46–93. [Google Scholar] [CrossRef] [Green Version]
  44. Nelyubina, Y.V.; Antipin, M.Y.; Lyssenko, K.A. Anion–anion interactions: Their nature, energy and role in crystal formation. Russ. Chem. Rev. 2010, 79, 167–187. [Google Scholar] [CrossRef]
  45. Dance, I.; Scudder, M. Supramolecular Motifs: Concerted Multiple Phenyl Embraces between Ph4P+ Cations are Attractive and Ubiquitous. Chem.-Eur. J. 1996, 2, 481–486. [Google Scholar] [CrossRef] [PubMed]
  46. Chesman, A.S.R.; Hodgson, J.L.; Izgorodina, E.I.; Urbatsch, A.; Turner, D.R.; Deacon, G.B.; Batten, S.R. Anion-anion interactions in the crystal packing of functionalized methanide anions: An experimental and computational study. Cryst. Growth Des. 2014, 14, 1922–1932. [Google Scholar] [CrossRef]
  47. Wysokiński, R.; Zierkiewicz, W.; Michalczyk, M.; Scheiner, S. Crystallographic and theoretical evidences of anion···anion interaction. ChemPhysChem 2021, 22, 818–821. [Google Scholar] [CrossRef]
  48. Nelyubina, Y.V.; Lyssenko, K.A.; Golovanov, D.G.; Antipin, M.Y. NO3···NO3 and NO3···π interactions in the crystal of urea nitrate. CrystEngComm 2007, 9, 991–996. [Google Scholar] [CrossRef]
  49. Khan, S.; Roy, S.; Harms, K.; Bauzá, A.; Frontera, A.; Chattopadhyay, S. Observation of an anion···anion interaction in a square planar copper(II) Schiff base complex: DFT study and CSD analysis. Inorg. Chim. Acta 2019, 487, 465–472. [Google Scholar] [CrossRef]
  50. Dougherty, D.A. Cation-pi interactions in chemistry and biology: A new view of benzene, Phe, Tyr, and Trp. Science 1996, 271, 163–168. [Google Scholar] [CrossRef]
  51. Ma, J.C.; Dougherty, D.A. The Cation-π Interaction. Chem. Rev. 1997, 97, 1303–1324. [Google Scholar] [CrossRef]
  52. Sussman, J.L.; Harel, M.; Frolow, F.; Oefner, C.; Goldman, A.; Tolker, L.; Silman, I. Atomic structure of acetylcholinesterase from Torpedo californica: A prototypic acetylcholine-binding protein. Science 1991, 253, 872–879. [Google Scholar] [CrossRef]
  53. Quiñonero, D.; Garau, C.; Rotger, C.; Frontera, A.; Ballester, P.; Costa, A.; Deyà, P.M. Anion-π Interactions: Do They Exist? Angew. Chem. 2002, 114, 3539–3542. [Google Scholar] [CrossRef]
  54. Schottel, B.L.; Chifotides, H.T.; Dunbar, K.R. Anion-π Interactions. Chem. Soc. Rev. 2008, 37, 68–83. [Google Scholar] [CrossRef] [PubMed]
  55. Wang, D.-X.; Wang, M.-X. Exploring Anion-π Interactions and Their Applications in Supramolecular Chemistry. Acc. Chem. Res. 2020, 53, 1364–1380. [Google Scholar] [CrossRef] [PubMed]
  56. Jo, H.J.; Jung, S.H.; Kim, H.-J. Synthesis and Hydrogen-Bonded Supramolecular Assembly of trans-Dihydroxotin(IV) Tetrapyridylporphyrin Complexes. Bull. Korean Chem. Soc. 2004, 25, 1869–1873. [Google Scholar]
  57. Jo, H.J.; Kim, S.H.; Kim, H.-J. Supramolecular Assembly of Tin(IV) Porphyrin Cations Stabilized by Ionic Hydrogen-Bonding Interactions. Bull. Korean Chem. Soc. 2015, 36, 2348–2351. [Google Scholar] [CrossRef]
  58. Tan, X.-W.; Li, H.-F.; Li, C.-H. Syntheses, Crystal Structures and Fluorescence Properties of Two Complexes Constructed from Rigid 1,4-bis(4-methyl-imidazolyl)Benzene. Chin. J. Inorg. Chem. 2017, 33, 156–162. [Google Scholar]
  59. Ni, T.-J.; Yuan, Q.-T.; Zhang, W.; Yang, Z.-J.; Yu, H.-H.; Cai, H.-R. Influence of Solvents on the Formation of Silver(I) Complexes with the Flexible 1,3,5-Tris(triazol-1-ylmethyl)-2,4,6-trimethylbenzene. Chin. J. Inorg. Chem. 2017, 33, 2177–2185. [Google Scholar]
  60. Bar, A.K.; Raghothoma, S.; Moon, D.; Mukherjee, P.S. Three-Component Self-Assembly of a Series of Triply Interlocked Pd12 Coordination Prisms and Their Non-Interlocked Pd6 Analogues. Chem. Eur. J. 2012, 18, 3199–3209. [Google Scholar] [CrossRef]
  61. Area Detector Control and Integration Software, Version 6.22; Bruker AXS Inc.: Madison, WI, USA, 1996.
  62. Sheldrick, G.M. SADABS: Empirical Absorption and Correction Software; University of Gottingen: Gottingen, Germany, 1999. [Google Scholar]
  63. Program for Solution and Refinement of Crystal Structures, Version 6.10; Bruker AXS Inc.: Madison, WI, USA, 2000.
Scheme 1. Preparation of SnP2+, and chemical structures of Sn(IV)-porphyrin complexes.
Scheme 1. Preparation of SnP2+, and chemical structures of Sn(IV)-porphyrin complexes.
Molbank 2022 m1454 sch001
Figure 1. Molecular structure of SnP2+. Ellipsoids were drawn at 50% probability, and hydrogen atoms of the porphyrin ligand were omitted for clarity.
Figure 1. Molecular structure of SnP2+. Ellipsoids were drawn at 50% probability, and hydrogen atoms of the porphyrin ligand were omitted for clarity.
Molbank 2022 m1454 g001
Figure 2. (a) Hydrogen bonds between the aqua ligands and the pyridyl substituents of SnP2+. (b) The two-dimensional array of the Sn(IV)-porphyrin cations via hydrogen bonding. Hydrogen bonds are indicated by dotted lines. Hydrogen atoms of the porphyrin ligand, nitrate anions, and clathrate acetone molecules were omitted for clarity.
Figure 2. (a) Hydrogen bonds between the aqua ligands and the pyridyl substituents of SnP2+. (b) The two-dimensional array of the Sn(IV)-porphyrin cations via hydrogen bonding. Hydrogen bonds are indicated by dotted lines. Hydrogen atoms of the porphyrin ligand, nitrate anions, and clathrate acetone molecules were omitted for clarity.
Molbank 2022 m1454 g002
Figure 3. (a) Position of nitrate anions in a confined space surrounded by Sn(IV)-porphyrin cations. (b) Hydrogen bonds between nitrate anions and aqua ligands. Hydrogen bonds are indicated by dotted lines. Hydrogen atoms of porphyrin ligand and clathrate acetone molecules were omitted for clarity.
Figure 3. (a) Position of nitrate anions in a confined space surrounded by Sn(IV)-porphyrin cations. (b) Hydrogen bonds between nitrate anions and aqua ligands. Hydrogen bonds are indicated by dotted lines. Hydrogen atoms of porphyrin ligand and clathrate acetone molecules were omitted for clarity.
Molbank 2022 m1454 g003
Figure 4. (a) Close contact between two nitrate anions. Dotted lines indicate anion···anion interactions. (b) Side view of the anion···π interaction indicated by dotted lines between the two centroids. (c) Top view along the axis connecting the two centroids. Pyridyl substituents and hydrogen atoms were omitted for clarity.
Figure 4. (a) Close contact between two nitrate anions. Dotted lines indicate anion···anion interactions. (b) Side view of the anion···π interaction indicated by dotted lines between the two centroids. (c) Top view along the axis connecting the two centroids. Pyridyl substituents and hydrogen atoms were omitted for clarity.
Molbank 2022 m1454 g004
Figure 5. (a) Packing view of alternating layers of Sn(IV)-porphyrin cations and nitrate anions. (b) Perspective view of the arrays. Hydrogen atoms and clathrate acetone molecules were omitted for clarity.
Figure 5. (a) Packing view of alternating layers of Sn(IV)-porphyrin cations and nitrate anions. (b) Perspective view of the arrays. Hydrogen atoms and clathrate acetone molecules were omitted for clarity.
Molbank 2022 m1454 g005
Table 1. Selected bond distances (Å) and angles (°) of SnP2+ in comparison with SnP6+ and SnP.
Table 1. Selected bond distances (Å) and angles (°) of SnP2+ in comparison with SnP6+ and SnP.
SnP2+SnP6+ [a]SnP [b]
Sn–O1W2.093(2)2.1024(19)2.014(3)
Sn–N12.093(2)2.087(2)2.098(3)
Sn–N22.089(2)2.078(2)2.104(3)
∠N1–Sn–N290.27(9)90.52(8)90.19(11)
∠N1–Sn–O1W90.95(9)89.64(8)93.12(11)
∠N2–Sn–O1W90.43(9)87.56(8)89.83(12)
Data from [a] [57] and [b] [56].
Table 2. Selected bond distances (Å) and angles (°) associated with hydrogen bonding (D, donor atom; A, acceptor atom).
Table 2. Selected bond distances (Å) and angles (°) associated with hydrogen bonding (D, donor atom; A, acceptor atom).
D–H···Ad(D–H)d(H···A)d(D···A)∠(DHA)
O1W–H1WA···N3i0.821.902.695(4)165.2
O1W–H1WB···O20.821.812.610(7)167.6
O1W–H1WB···N50.822.623.394(7)157.8
Symmetry codes: (i) −x+1, y+1/2, −z+1/2.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Kim, H.-J. Assembly of Sn(IV)-Porphyrin Cation Exhibiting Supramolecular Interactions of Anion···Anion and Anion···π Systems. Molbank 2022, 2022, M1454. https://doi.org/10.3390/M1454

AMA Style

Kim H-J. Assembly of Sn(IV)-Porphyrin Cation Exhibiting Supramolecular Interactions of Anion···Anion and Anion···π Systems. Molbank. 2022; 2022(4):M1454. https://doi.org/10.3390/M1454

Chicago/Turabian Style

Kim, Hee-Joon. 2022. "Assembly of Sn(IV)-Porphyrin Cation Exhibiting Supramolecular Interactions of Anion···Anion and Anion···π Systems" Molbank 2022, no. 4: M1454. https://doi.org/10.3390/M1454

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop