Next Article in Journal
Sol-Gel Synthesis and Characterization of Highly Selective Poly(N-methyl pyrrole) Stannous(II)Tungstate Nano Composite for Mercury (Hg(II)) Detection
Previous Article in Journal
Mechanochemical Synthesis of Nanocrystalline Olivine-Type Mg2SiO4 and MgCoSiO4
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Glycols in the Synthesis of Zinc-Anilato Coordination Polymers

by
Olesya Y. Trofimova
1,
Arina V. Maleeva
1,
Kseniya V. Arsenyeva
1,
Anastasiya V. Klimashevskaya
1,
Il’ya A. Yakushev
2,3 and
Alexandr V. Piskunov
1,*
1
Laboratory of Metal Complexes with Redox-Active Ligands, G.A. Razuvaev Institute of Organometallic Chemistry, Russian Academy of Sciences, 603137 Nizhny Novgorod, Russia
2
N.S. Kurnakov Institute of General and Inorganic Chemistry, Russian Academy of Sciences, 119991 Moscow, Russia
3
Kurchatov Institute, National Research Center, 123182 Moscow, Russia
*
Author to whom correspondence should be addressed.
Crystals 2022, 12(3), 370; https://doi.org/10.3390/cryst12030370
Submission received: 27 February 2022 / Revised: 7 March 2022 / Accepted: 8 March 2022 / Published: 10 March 2022

Abstract

:
We report the synthesis, structural investigation, and thermal behavior for three zinc-based 1D-coordination polymers with 3,6-di-tert-butyl-2,5-dihydroxy-p-benzoquinone, which were synthesized in the presence of different glycols. The interaction of zinc nitrate with glycols, followed by using the resulting solution in solvothermal synthesis with the anilate ligand in DMF, makes it possible to obtain linear polymer structures with 1,2-ethylene or 1,2-propylene glycols coordinated to the metal. The reaction involving 1,3-propylene glycol under similar conditions gives a crystal structure that does not contain a diol. The crystal and molecular structures of the synthesized compounds were determined using single crystal by X-ray structural analysis. The influence of glycol molecules coordinated to the metal on the thermal destruction of synthesized compounds is shown.

1. Introduction

The development of synthetic approaches to produce new metal-organic frameworks (MOF) is one of the perspective directions for modern supramolecular chemistry. The possibility of structural aimed design and influence on the final property of coordination polymer by choosing a metal center and organic linkers opens up broad prospects for obtaining multifunctional materials [1]. MOFs have an extensive set of properties that ensure their potential application in various fields of both fundamental and applied science. Materials based on coordination polymers can be used as gas adsorbents [2,3,4], electrochemical or photophysical sensors [5], as well as luminescent [6,7], photocatalytic [8,9], electrically conductive [10,11], and magnetic [12,13] materials. Metal-organic redox-active coordination polymers represent a particular class of derivative [14,15,16,17]. The presence of a redox-active bridging ligand in the MOF gives prospects for the reversible switching of polymer properties during its oxidation/reduction without destruction of the framework [18,19,20,21]. Anilate redox-active ligands are derivatives of 2,5-dihydroxy-para-benzoquinone with various substituents in the 3,6 positions, which affect the electronic structure of the quinoid ring (H, F, Cl, Br, CN, NO2, t-Bu, etc.). It has a direct effect on the properties of MOFs based on them [22]. In this regard, the study of new ancillary ligands with various substituents in the quinoid ring is one of the promising tasks for expanding the range of properties of redox-active MOFs [4,23,24]. The study of 2,5-dihydroxy-3,6-di-tert-butyl-para-quinone (H2pQ) [25] as a ditopic ligand in the synthesis of metal-organic frameworks started recently [26,27,28]. Previously, mononuclear derivatives of triphenylantimony(V) [29] and rare binuclear complexes of tin, nickel, iron, and cobalt with this ligand were already known [25,30,31].
Anilate ligands bound to metals can exist in four redox states, as presented in Figure 1. In most derivatives formed by anilate ligands, these linkers are in the dianionic state.
The introduction of additional linkers into the metal-organic ligand system is essential in the design of MOFs. It changes the polymer structure and affects the physicochemical properties of the resulting derivative. For example, intensive research has begun on the properties of heteroleptic derivatives containing simultaneously two types of ligands in the MOF: anilate and dicarboxylate [27,32,33]. The recent works [34,35] demonstrated the use of polyatomic alcohols as anionic ligands in the synthesis of heteroleptic zinc MOFs. These coordination polymers are mesoporous and exhibit unique sorption behavior.
The present paper reports on new linear zinc-based coordination polymers with H2pQ. The effect of the glycol ligand in the metal coordination sphere on the MOFs’ structural properties and thermal stability is considered.

2. Materials and Methods

2.1. Reagents and Methods

All commercial reagents were purchased from Sigma Aldrich (Darmstadt, Germany) and used without purification. 2,5-dihydroxy-3,6-di-tert-butyl-p-benzoquinone(H2pQ) was prepared following the previously published methodology [25]. All synthetic manipulations were performed under conditions exluding air oxygen. Solvents were purified using standard techniques [36]. CHN-analysis was performed with an Elementar Vario El cube equipment (Elementar Analysensysteme GmbH, Langenselbold, Germany). TGAs of compounds were examined on a Mettler Toledo TGA/DSC3+ instrument (Columbus, OH, USA) from 30 to 500 °C with a PCA pan and heat rate of 5 °C/min under a N2 atmosphere.

2.2. Synthesis of [Zn(pQ)EG]·2DMF (1), [Zn(pQ)1,2-PG]·2DMF (2) and [Zn(pQ)(DMF)2] (3)

Zn(NO3)2·6H2O (0.04 mmol) was dissolved in 3 mL of respective glycol (ethylene glycol (EG) for 1, 1,2-propylene (1,2-PG) glycol for 2 and 1,3-propylene glycol (1,3-PG) for 3) and heated under stirring at 100 °C for 1 h. The solution of H2pQ (0.04 mmol) in 3 mL of N,N-dimethylformamide (DMF) was added. The resulted reaction mixture was kept at 50 °C in a glass vial. The formation of crystals of 1, 2, and 3 was observed after 4, 15, and 5 days, respectively. Bright-colored burgundy crystals of 1–3 were filtered, washed by DMF, and dried in the air.
[Zn(pQ)EG]·2DMF (1). Yield 70%. Elemental analysis calculated for C22H38N2O8Zn (%): C, 50.43; H, 7.31; N, 5.35. Found (%): C, 49.97; H, 7.64; N 5.21. IR (Nujol, KBr, cm–1): 3200w (large), 1652w, 1417w, 1391w, 1341w, 1258w, 1214w, 1204w, 1102w, 1077w, 977m, 911w, 890w, 864m, 793w, 673w, 660w, 627m, 564m, 529w, 507w.
[Zn(pQ)1,2-PG]·2DMF (2). Yield 49%. Elemental analysis calculated for C23H40N2O8Zn (%): C 51.35; H 7.49; N 5.21. Found (%): C 51.34; H 7.76; N 5.29. IR (Nujol, KBr, cm–1): 3300w (large), 1660w, 1534w, 1501w, 1409w, 1288w, 1257w, 1221w, 1102w, 1078w, 1039w, 998m, 911w, 896m, 877m, 844w, 779m, 667w, 591w, 578m, 554m.
[Zn(pQ)(DMF)2] (3). Yield 68%. Elemental analysis calculated for C22H36N2O6Zn (%): C 52.01; H 6.98; N 6.07. Found (%): C 51.93; H 7.10; N 5.92. IR (Nujol, KBr, cm–1): 1530w, 1345w, 1259w, 1217w, 1198w, 1110w, 1063m, 1049m, 1012m, 971m, 928m, 904w, 867m, 791w, 681w, 658w, 624m, 529s, 503s.

2.3. Single-Crystal X-ray Diffraction Studies

The single-crystal X-ray analysis for the complexes 1 and 3 was performed using a Bruker D8 Venture Photon single-crystal diffractometer (Billerica, MA, USA) equipped with microfocus sealed tube Incoatec IµS 3.0 (Mo radiation, λ = 0.71073 Å) in φ- and ω-scan mode at the center of shared equipment IGIC RAS (Moscow, Russia). X-ray diffraction data for the coordination polymer 2 were obtained on the “Belok” beamline at the Kurchatov Synchrotron Radiation Source (National Research Center “Kurchatov Institute”, Moscow, Russia) in φ-scan mode using SX165 CCD detector (Rayonix, Evanston, IL, USA), λ = 74,500 Å [37].
The raw data for 1 and 3 were processed with the APEX3 software (version 2016.9-0; Bruker, 2016) [38]; experimental intensities were revised for absorption effects using SADABS (version 2016/2; Bruker, 2016) [38] and TWINABS (version 2012/1; Bruker, 2012) programs [39] for 1 and 3, respectively. The raw data for 2 were processed using the XDS data reduction program (version Feb. 5, 2021) [40], comprising absorption correction.
The crystal structures were solved by direct methods [41] and refined by the full-matrix least-squares on F2 [42] using OLEX2 structural data visualization and analysis program [43]. The twinned data for 2 were treated using PLATON software (version 30814) [44] followed by refinement using HKLF 5 instruction.
All non-hydrogen atoms were refined, applying anisotropic displacement parameters without any constraints or restraints for 13, except disordered over two-position (occupancy ratio 0.85(1)/0.15(1)) tert-butyl moiety in the structure 1, which was refined using geometrical restraints (SADI); carbon atoms related to the minor part of the disordered position were refined isotropically. In the general case, the hydrogen atoms were placed in the ideal calculated positions and refined using the riding model with Uiso(H) = 1.5Ueq(C) for the methyl groups and with Uiso(H) = 1.2Ueq(C) for other hydrogen atoms, while H-atoms of the alcohol groups were found from difference Fourier synthesis and refined with isotropic thermal parameters with restrained O-H distances (SADI, DFIX).
Oxford Cryosystems Cryostream 800 (Long Hanborough, UK) open-system cooler instrument was used to performing low-temperature X-ray diffraction experiments. Table 1 summarizes data of crystal data and structure refinement details for 13. CIF and CheckCIF files for 13 can be found at Supplementary Materials.

3. Results and Discussions

Zinc cations have proven to be a reasonably convenient platform for testing methods for obtaining coordination polymers based on anilate ligands [3,45,46,47,48,49,50,51,52,53]. The first examples of such derivatives, 13, with a tert-butyl-substituted ligand were obtained in this work. In addition, attempts have been made to introduce glycols into the composition of the resulting coordination polymer. The synthesis of zinc derivatives includes two stages (Figure 2). The glycol ligand is coordinated to the metal atom at the first stage. Then, a solution of H2pQ in DMF was added, and a slow reaction of formation of linear polymers 1–3 proceeded. It is worth noting that the reaction including 1,3-propylene glycol at the first stage gives coordination polymer 3, which does not contain the diol in the resulting product. This result points to the weak coordination of 1,3-propylene glycol. This diol is displaced by a strong coordinating solvent (DMF) from the coordination sphere of zinc.
Derivatives 13 are isolated from the reaction mixture as burgundy crystals. The synthesized MOFs are analytically pure after washing on a filter with DMF and further drying in air. Coordination polymers 1 and 2 contain two DMF molecules per one unit of polymer chain according to the elemental analysis. Compounds 13 are completely insoluble in all organic solvents and in water. They are are stable to oxygen and atmospheric moisture.
The single-crystal X-ray diffraction determined the crystal and molecular structure of 13. Thermogravimetric and elemental analysis methods have confirmed the composition and chemical purity of all compounds. Molecular structures of 13 are shown in Figure 3, Figure 4 and Figure 5, respectively. The picked bond lengths are presented in Table 2.
According to the X-ray data, complexes 13 are linear zigzag polymers (Figure 3, Figure 4 and Figure 5). Crystal of compound 1 is characterized by the monoclinic symmetry group C2/c, and derivatives 2 and 3 crystallize in the triclinic symmetry group P-1. The coordination environment of each of the zinc cations in compounds 13 is a distorted octahedron, along the vertices of which there are four oxygens of two dianionic ligands pQ2− and two oxygens of O-donor neutral ligands—EG for 1; 1,2-PG for 2; and two DMF for 3 (Figure 3, Figure 4 and Figure 5). Cis-location of ligands in the metal coordination sphere leads to a zigzag structure of polymer chains.
The pQ dianion in 13 has a similar structure. It demonstrates the presence of two unbonded anionic π-electron systems—OCCCO. These systems are interconnected by single bonds C(2)–C(4B) (for 1), C(4)–C(6A) and C(11)–C(13B) (for 2), and C(1)–C(3A) and C(8)–C(10B) (for 3). The element of symmetry passes through the middle of these bonds. The C(4)–C(5) and C(5)–C(6) bonds, as well as C(4)–O(3) and C(6A)–O(4A) bonds are near equal to each other in compound 1. It indicates a high delocalization of the charge in these ligand fragments. The lengths of the C(4)–C(5) and C(5)–C(6) bonds are in the range which is normal for aromatic carbon–carbon bonds. The interatomic distances C(4)-O(3) and C(6A)-O(4A) have average values between those for double and single carbon–oxygen bonds. Derivatives 2 and 3 are characterized by the same charge distribution in the organic ligand. Such electron density delocalization is usual for bridging anilate ligands in dinuclear metal complexes and coordination polymers. The above-mentioned bond length distribution in the dianion pQ2− was recently observed in the related magnesium derivative [28] and differs sharply from that observed for the mononuclear triphenylantimony(V) complex [29], for which an exact o-quinoid type of bond length alternation was found.
The lengths of Zn–O bonds with p-quinoid ligands in 1–3 are practically aligned and are in the range 2.04–2.10 Å. They are less than the sum of the covalent radii of the corresponding elements (2.24 Å) [54]. The coordination of neutral glycolic fragments (EG in 1 and 1,2-PG in 2) is characterized by zinc–oxygen bond lengths in the range 2.13–2.14 Å, which is typical for this type of coordination [55,56,57]. Zinc–oxygen bond lengths with DMF in 3 are quite similar to Zn–O bonds with pQ-ligands. It points to the strong coordination bonds, which are destroyed at sufficiently high temperatures, as shown by thermogravimetric analysis.
The glycol-containing compounds 1 and 2 have a free volume of 32.5% and 37.5% available for solvent, respectively, occupied by two guest DMF molecules per polymer unit in both cases (Figure 6).
The formation of voids in crystals of 1 and 2 is caused by strong hydrogen bonds. They form between the OH-hydrogens of the glycol ligand and the guest DMF molecules’ oxygen atoms (Figure 7).
The thermal stability of 13 was studied by the TGA (Figure 8). Studies have shown that the thermal behavior of complexes 1 and 2 containing a glycol ligand in the metal coordination sphere is very similar. However, they are quite different from the thermal decomposition of derivative 3 containing coordinated DMF molecules.
Coordination polymers [Zn(pQ)EG)]·2DMF (1) and [Zn(pQ)1,2-PG)]·2DMF (2), according to the TGA data, contain an occluded DMF molecules. They are removed from the crystals at a temperature of 30–70 °C (mass loss is 2 and 2.5% for 1 and 2, respectively). The guest DMF molecules from the voids of 1 and 2 crystals are removed at 70–150 °C. The mass loss is 28% for both samples. It is in good agreement with the content of two DMF molecules per unit of respective coordination polymers. The decoordination of glycol molecules occurs upon further heating from 170 to 300 °C. The transformation of the anilate ligand accompanies this process. At this stage, the total mass loss is 31% and 32% for 1 and 2, respectively. It is responsible, along with the loss of glycols, for the destruction of tert-butyl substituents, which are removed in the form of two isobutylene molecules. The subsequent increase in temperature up to 500 °C leads to the gradual destruction of the polymer containing the dealkylated anilate ligand. [Zn(pQ)(DMF)2] (3) does not contain occluded and guest solvents. Therefore, the first stage of mass loss (16%) in the temperature range of 80–140 °C corresponds to the decoordination of one DMF molecule from a zinc atom. Heating up to 230 °C is accompanied by the cleavage of the second DMF molecule (mass loss is 15%). Further, in the temperature range of 360–400 °C, the final decomposition of the polymer occurs, at which the destruction of the anilate ligand occurs. The mass loss at this stage of TGA is 36%.
We should note that the thermal stability of zinc complexes significantly exceeds that of similar magnesium derivatives [28]. Meanwhile, the thermal behavior of compound 3 is comparable to the data obtained for 2D grids based on lanthanide ions [26] with the same ligands. When comparing glycol-containing polymers 1 and 2 with complex 3, it is unambiguously possible to conclude that the hydroxyl-containing ligand promotes the dealkylation of the anilate ligand at sufficiently low temperatures, which was not observed in other studied derivatives based on this para-quinone ligand [26,27,28].

4. Conclusions

Thus, 1D metal-organic coordination polymers of zinc were synthesized based on the 2,5-dihydroxy-3,6-di-tert-butyl-para-benzoquinone. It is shown that the introduction of glycol ligands into the coordination sphere of the metal contributes to the formation of porous structures. According to the TGA data, the presence of glycols in zinc derivatives promotes the dealkylation of anilate ligands in the temperature range of 170–300 °C, with further slow destruction of the polymer. Contrastingly, for the compound, which does not contain a diol in the metal coordination sphere, the destruction of the organic ligand occurs in a narrow range of high temperatures (360–400 °C).

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/cryst12030370/s1, CIF and CheckCIF files for 13. Crystallographic data for the structural analysis have been deposited with the Cambridge Crystallographic Data Centre, CCDC No. 2153350 (1), 2153351 (2), 2153352 (3). Copies of the above information may be received free of charge from The Director, CCDC, 12, Union Road, Cambridge CB2 1EZ, U.K.; fax +44-1223-336033; e-mail [email protected]

Author Contributions

Supervision, review and editing, conceptualization, A.V.P.; writing—original draft preparation, O.Y.T.; discussion, synthesis and investigation, O.Y.T., A.V.K.; synthesis, A.V.M., K.V.A.; discussion and formal analysis, I.A.Y. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by Russian Science Foundation, grant number 22-23-00750.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data presented in this study are available in the article and Supplementary Materials.

Acknowledgments

The studies were carried out using the equipment of the center for collective use “Analytical Center of the IOMC RAS” with the financial support of the grant “Ensuring the development of the material and technical infrastructure of the centers for collective use of scientific equipment” (unique identifier RF—2296.61321x0017; Agreement Number—075-15-2021-670).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Agafonov, M.A.; Alexandrov, E.V.; Artyukhova, N.A.; Bekmukhamedov, G.E.; Blatov, V.A.; Butova, V.V.; Gayfulin, Y.M.; Gharibyan, A.A.; Gafurov, Z.N. Metal-organic coordination polymers in Russia: From synthesis and structure to functional properties and materials. J. Struct. Chem. 2022. [Google Scholar] [CrossRef]
  2. Kingsbury, C.J.; Abrahams, B.F.; Auckett, J.E.; Chevreau, H.; Dharma, A.D.; Duyker, S.; He, Q.; Hua, C.; Hudson, T.A.; Murray, K.S.; et al. Square Grid Metal–Chloranilate Networks as Robust HostSystems for GuestSorption. Chem. Eur. J. 2019, 25, 5222–5234. [Google Scholar] [CrossRef] [PubMed]
  3. Abrahams, B.F.; Dharma, A.D.; Dyett, B.; Hudson, T.A.; Maynard-Casely, H.; Kingsbury, C.J.; McCormick, L.J.; Robson, R.; Sutton, A.L.; White, K.F. An indirect generation of 1D M(II)-2,5-dihydroxybenzoquinone coordination polymers, their structural rearrangements and generation of materials with a high affinity for H2, CO2 and CH4. Dalton Trans. 2016, 45, 1339–1344. [Google Scholar] [CrossRef] [PubMed]
  4. Monni, N.; Andres-Garcia, E.; Caamaño, K.; García-López, V.; Clemente-Juan, J.M.; Giménez-Marqués, M.; Oggianu, M.; Cadoni, E.; Espallargas, G.M.; Clemente-León, M.; et al. A thermally/chemically robust and easily regenerable anilatobased ultramicroporous 3D MOF for CO2 uptake and separation. J. Mater. Chem. A 2021, 9, 25189–25195. [Google Scholar] [CrossRef]
  5. Fasna, P.H.F.; Sasi, S. A Comprehensive Overview on Advanced Sensing Applications of Functional Metal Organic Frameworks (MOFs). ChemistrySelect 2021, 6, 6365–6379. [Google Scholar] [CrossRef]
  6. Huangfu, M.; Wang, M.; Lin, C.; Wang, J.; Wu, P. Luminescent metal–organic frameworks as chemical sensors based on “mechanism–response”: A review. Dalton Trans. 2021, 50, 3429–3449. [Google Scholar] [CrossRef] [PubMed]
  7. Li, P.; Zhou, Z.; Zhao, Y.S.; Yan, Y. Recent advances in luminescent metal-organic frameworks and their photonic applications. Chem. Comm. 2021, 57, 13678–13691. [Google Scholar] [CrossRef] [PubMed]
  8. Ezugwu, C.I.; Liu, S.; Li, C.; Zhuiykov, S.; Roy, S.; Verpoort, F. Engineering metal-organic frameworks for efficient photocatalytic conversion of CO2 into solar fuels. Coord. Chem. Rev. 2021, 450, 214245. [Google Scholar] [CrossRef]
  9. Zhang, X.; Wang, C.; Wang, L.Y.; Xing, Y.H.; Bai, F.Y.; Shi, Z. Construction of crystalline cadmium complex based on 1,4,5,8-naphthalene diimide derivative and photocatalytic degradation about organic dyes. Appl. Organomet. Chem. 2022, e6603, 1–13. [Google Scholar] [CrossRef]
  10. Wang, M.; Dong, R.; Feng, X. Two-dimensional conjugated metal–organic frameworks (2D c-MOFs): Chemistry and function for MOFtronics. Chem. Soc. Rev. 2021, 50, 2764–2793. [Google Scholar] [CrossRef] [PubMed]
  11. Dong, R.; Feng, X. Making large single crystals of 2D MOFs. Nat. Mater. 2021, 20, 122–131. [Google Scholar] [CrossRef] [PubMed]
  12. Monni, N.; Oggianu, M.; Sahadevan, S.A.; Mercuri, M.L. Redox Activity as a Powerful Strategy to Tune Magnetic and/or Conducting Properties in Benzoquinone-Based Metal-Organic Frameworks. Magnetochemistry 2021, 7, 109. [Google Scholar] [CrossRef]
  13. Espallargas, G.M.; Coronado, E. Magnetic functionalities in MOFs: From the framework to the pore. Chem. Soc. Rev. 2018, 47, 533–557. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  14. Huang, Z.; Yu, H.; Wang, L.; Liu, X.; Lin, T.; Haq, F.; Vatsadze, S.Z.; Lemenovskiy, D.A. Ferrocene-contained metal organic frameworks: From synthesis to applications. Coord. Chem. Rev. 2021, 430, 213737. [Google Scholar] [CrossRef]
  15. Calbo, J.; Golomb, M.J.; Walsh, A. Redox-active metal–organic frameworks for energy conversion and storage. J. Mater. Chem. A 2019, 7, 16571–16597. [Google Scholar] [CrossRef]
  16. D’Alessandro, D.M. Exploiting redox activity in metal–organic frameworks: Concepts, trends and perspectives. Chem. Comm. 2016, 52, 8957–8971. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  17. Kitagawa, S.; Kawata, S. Coordination compounds of 1,4-dihydroxybenzoquinone and its homologues. Structures and properties. Coord. Chem. Rev. 2002, 224, 11–34. [Google Scholar] [CrossRef]
  18. Chang, C.-H.; Li, A.-C.; Popovs, I.; Kaveevivitchai, W.; Chen, J.-L.; Chou, K.-C.; Kuof, T.-S.; Chen, T.-H. Elucidating metal and ligand redox activities of a copper-benzoquinoid coordination polymer as the cathode for lithium-ion batteries. J. Mater. Chem. A 2019, 7, 23770–23774. [Google Scholar] [CrossRef]
  19. Abrahams, B.F.; Bond, A.M.; Le, T.H.; McCormick, L.J.; Nafady, A.; Robson, R.; Vo, N. Voltammetric reduction and re-oxidation of solid coordination polymers of dihydroxybenzoquinone. Chem. Comm. 2012, 48, 11422–11424. [Google Scholar] [CrossRef] [PubMed]
  20. DeGayner, J.A.; Jeon, I.-R.; Sun, L.; Dinca, M.; Harris, T.D. 2D Conductive Iron-Quinoid Magnets Ordering up to T= 105 K via Heterogenous Redox Chemistry. J. Am. Chem. Soc. 2017, 139, 4175–4184. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  21. Monni, N.; Angotzi, M.S.; Oggianu, A.M.; Sahadevan, S.A.; Mercuri, M.L. Redox-active benzoquinones as challenging ‘‘non-innocent’’ linkers to construct 2D frameworks and nanostructures with tunable physical properties. J. Mater. Chem. C 2022, 10, 1548–1572. [Google Scholar] [CrossRef]
  22. Mercuri, M.L.; Congiu, F.; Concas, G.; Sahadevan, S.A. Recent Advances on Anilato-Based Molecular Materials with Magnetic and/or Conducting Properties. Magnetochemistry 2017, 3, 17. [Google Scholar] [CrossRef] [Green Version]
  23. Milašinović, V.; Molčanov, K. Nitranilic acid as a basis for construction of coordination polymers: From discrete monomers to 3D networks. CrystEngComm 2019, 21, 2962–2969. [Google Scholar] [CrossRef]
  24. Milašinović, V.; Jurić, M.; Molčanov, K. Nitrochloranilic acid: A novel asymmetrically substituted quinoid bridging ligand for design of coordination polymers. CrystEngComm 2021, 23, 2304–2315. [Google Scholar] [CrossRef]
  25. Khamaletdinova, N.M.; Meshcheryakova, I.N.; Piskunov, A.V.; Kuznetsova, O.V. Experimental and theoretical study of the vibrational spectra of tin(IV) complexes based on 2-hydroxy-3,6-di-tert-butyl-para-benzoquianone. J. Struct. Chem. 2015, 56, 233–242. [Google Scholar] [CrossRef]
  26. Kharitonov, A.D.; Trofimova, O.Y.; Meshcheryakova, I.N.; Fukin, G.K.; Khrizanforov, M.N.; Budnikova, Y.H.; Bogomyakov, A.S.; Aysin, R.R.; Kovalenko, K.A.; Piskunov, A.V. 2D-Metal-organic coordination polymers of lanthanides (La(III), Pr(III) and Nd(III)) with redox-active dioxolene bridging ligand. CrystEngComm 2020, 22, 4675–4679. [Google Scholar] [CrossRef]
  27. Trofimova, O.Y.; Maleeva, A.V.; Ershova, I.V.; Cherkasov, A.V.; Fukin, G.K.; Aysin, R.R.; Kovalenko, K.A.; Piskunov, A.V. Heteroleptic LaIII Anilate/Dicarboxylate Based Neutral 3D-Coordination Polymers. Molecules 2021, 26, 2486. [Google Scholar] [CrossRef] [PubMed]
  28. Trofimova, O.Y.; Ershova, I.V.; Maleeva, A.V.; Yakushev, I.A.; Dorovatovskii, P.V.; Aisin, R.R. Metal–organic frameworks of magnesium based on 2,5-dihydroxy-3,6-di-tert-butyl-para-benzoquinone. Russ. J. Coord. Chem. 2021, 47, 610–619. [Google Scholar] [CrossRef]
  29. Okhlopkova, L.S.; Poddel’sky, A.I.; Fukin, G.K.; Smolyaninov, I.V. Triphenylantimony(v) catecholates based on 3,6-di-tert-butyl-2,5-dihydroxy-1,4-benzoquinone. Russ. J. Coord. Chem. 2020, 46, 386–393. [Google Scholar] [CrossRef]
  30. Min, K.S.; DiPasquale, A.G.; Rheingold, A.L.; White, H.S.; Miller, J.S. Observation of Redox-Induced Electron Transfer and Spin Crossover for Dinuclear Cobalt and Iron Complexes with the 2,5-Di-tert-butyl-3,6-dihydroxy-1,4-benzoquinonate Bridging Ligand. J. Am. Chem. Soc. 2009, 131, 6229–6236. [Google Scholar] [CrossRef] [PubMed]
  31. Min, K.S.; DiPasquale, A.; Rheingold, A.L.; Miller, J.S. Room-Temperature Spin Crossover Observed for [(TPyA)FeII(DBQ2-)FeII(TPyA)]2+ [TPyA=Tris(2-pyridylmethyl)amine; DBQ2-=2,5-Di-tert-butyl-3,6-dihydroxy-1,4-benzoquinonate]. Inorg. Chem. Commun. 2007, 46, 1048–1050. [Google Scholar] [CrossRef] [PubMed]
  32. Sahadevan, S.A.; Monni, N.; Oggianu, M.; Abhervé, A.; Marongiu, D.; Saba, M.; Mura, A.; Bongiovanni, G.; Mameli, V.; Cannas, C.; et al. Heteroleptic NIR-Emitting YbIII/Anilate-Based Neutral Coordination Polymer Nanosheets for Solvent Sensing. ACS Appl. Nano Mater. 2020, 3, 94–104. [Google Scholar] [CrossRef] [Green Version]
  33. Sahadevan, S.A.; Manna, F.; Abhervé, A.; Oggianu, M.; Monni, N.; Mameli, V.; Marongiu, D.; Quochi, F.; Gendron, F.; Guennic, B.L.; et al. Combined Experimental/Theoretical Study on the Luminescent Properties of Homoleptic/Heteroleptic Erbium(III) Anilate-Based 2D Coordination Polymers. Inorg. Chem. 2021, 60, 17765–17774. [Google Scholar] [CrossRef] [PubMed]
  34. Lysova, A.A.; Samsonenko, D.G.; Dorovatovskii, P.V.; Lazarenko, V.A.; Khrustalev, V.N.; Kovalenko, K.A.; Dybtsev, D.N.; Fedin, V.P. Tuning the Molecular and Cationic Affinity in a Series of Multifunctional Metal−Organic Frameworks Based on Dodecanuclear Zn(II) Carboxylate Wheels. J. Am. Chem. Soc. 2019, 141, 17260–17269. [Google Scholar] [CrossRef] [PubMed]
  35. Lysova, A.A.; Samsonenko, D.G.; Kovalenko, K.A.; Nizovtsev, A.S.; Dybtsev, D.N.; Fedin, V.P. A Series of Mesoporous Metal-Organic Frameworks with Tunable Windows Sizes and Exceptionally High Ethane over Ethylene Adsorption Selectivity. Angew. Chem. Int. Ed. 2020, 59, 20561–20567. [Google Scholar] [CrossRef] [PubMed]
  36. Perrin, D.D.; Armarego, W.L.F.; Perrin, D.R. Purification of Laboratory Chemicals; Elsevier: Oxford, UK, 1980. [Google Scholar]
  37. Svetogorov, R.D.; Dorovatovskii, P.V.; Lazarenko, V.A. Belok/XSA diffraction beamline for studying crystalline samples at Kurchatov synchrotron radiation source. Cryst. Res. Technol. 2020, 55, 1900184. [Google Scholar] [CrossRef]
  38. APEX3, SAINT and SADABS; Bruker AXS Inc.: Madison, WI, USA, 2016.
  39. Sheldrick, G.M. TWINABS, version 2012/1; Bruker AXS Inc.: Madison, WI, USA, 2012. [Google Scholar]
  40. Kabsch, W. XDS. Acta Crystallogr. D 2010, 66, 125–132. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  41. Sheldrick, G.M. SHELXT—Integrated space group and crystal structure determination. Acta Crystallogr. A 2015, 71, 3–8. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  42. Sheldrick, G.M. Crystal structure refinement with SHELXL. Acta Crystallogr. C 2015, 71, 3–8. [Google Scholar] [CrossRef] [PubMed]
  43. Dolomanov, O.V.; Bourhis, L.J.; Gildea, R.J.; Howard, J.A.K.; Puschmann, H. OLEX2: A complete structure solution, refinement and analysis program. J. Appl. Crystallogr. 2009, 42, 339–341. [Google Scholar] [CrossRef]
  44. Spek, A.L. Single-crystal structure validation with the program PLATON. J. Appl. Crystallogr. 2003, 36, 7–11. [Google Scholar] [CrossRef] [Green Version]
  45. Han, Y.; Li, X.; Li, L.; Ma, C.; Shen, Z.; Song, Y.; You, X. Structures and Properties of Porous Coordination Polymers Based on Lanthanide Carboxylate Building Units. Inorg. Chem. 2010, 49, 10781–10787. [Google Scholar] [CrossRef] [PubMed]
  46. Liu, L.; Li, L.; Ziebel, M.E.; Harris, T.D. Metal−Diamidobenzoquinone Frameworks via Post-Synthetic Linker Exchange. J. Am. Chem. Soc. 2020, 142, 4705–4713. [Google Scholar] [CrossRef] [PubMed]
  47. Liu, L.; Li, L.; DeGayner, J.A.; Winegar, P.H.; Fang, Y.; Harris, T.D. Harnessing Structural Dynamics in a 2D Manganese−Benzoquinoid Framework To Dramatically Accelerate Metal Transport in Diffusion-Limited Metal Exchange Reactions. J. Am. Chem. Soc. 2018, 140, 11444–11453. [Google Scholar] [CrossRef] [PubMed]
  48. Abrahams, B.F.; Hudson, T.A.; McCormick, L.J.; Robson, R. Coordination Polymers of 2,5-Dihydroxybenzoquinone and Chloranilic Acid with the (10,3)-a Topology. Cryst. Growth Des. 2011, 11, 2717–2720. [Google Scholar] [CrossRef]
  49. Abrahams, B.F.; Lu, K.D.; Moubaraki, B.; Murray, K.S.; Robson, R. X-Ray diffraction and magnetic studies on a series of isostructural divalent metal chloranilates with zigzag polymeric chain structures and on a dinuclear iron(III) chloranilate. Dalton Trans. 2000, 1793–1797. [Google Scholar] [CrossRef]
  50. Luo, T.-T.; Liu, Y.-H.; Tsai, H.-L.; Su, C.-C.; Ueng, C.-H.; Lu, K.-L. A Novel Hybrid Supramolecular Network Assembled from Perfect π-π Stacking of an Anionic Inorganic Layer and a Cationic Hydronium-Ion- Mediated Organic Layer. Eur. J. Inorg. Chem. 2004, 2004, 4253–4258. [Google Scholar] [CrossRef]
  51. Kingsbury, C.J.; Abrahams, B.F.; D’Alessandro, D.M.; Hudson, T.A.; Murase, R.; Robson, R.; White, K.F. Role of NEt4+ in Orienting and Locking Together [M2lig3]2− (6,3) Sheets (H2lig = Chloranilic or Fluoranilic Acid) to Generate Spacious Channels Perpendicular to the Sheets. Cryst. Growth Des. 2017, 17, 1465–1470. [Google Scholar] [CrossRef]
  52. Jeon, I.-R.; Negru, B.; Duyne, R.P.V.; Harris, D. A 2D Semiquinone Radical-Containing Microporous Magnet with Solvent-Induced Switching from Tc = 26 to 80 K. J. Am. Chem. Soc. 2015, 137, 15699–15702. [Google Scholar] [CrossRef] [PubMed]
  53. Murase, R.; Abrahams, B.F.; D’Alessandro, D.M.; Davies, C.G.; Hudson, T.A.; Jameson, G.N.L.; Moubaraki, B.; Murray, K.S.; Robson, R.; Sutton, A.L. Mixed Valency in a 3D Semiconducting Iron−Fluoranilate Coordination Polymer. Inorg. Chem. 2017, 56, 9025–9035. [Google Scholar] [CrossRef] [PubMed]
  54. Batsanov, S.S. The atomic radii of the elements. Russ. J. Inorg. Chem. 1991, 36, 1694–1705. [Google Scholar]
  55. Song, T.-Q.; Dong, J.; Gao, H.-L.; Cui, J.-Z. Three coordination polymers based on M2 (M = Co, Ni and Zn) clusters: Structures, magnetic and fluorescent properties. Inorg. Chim. Acta 2017, 466, 393–397. [Google Scholar] [CrossRef]
  56. Cheplakova, A.M.; Kovalenko, K.A.; Samsonenko, D.G.; Vinogradov, A.S.; Karpov, V.M.; Platonovc, V.E.; Fedin, V.P. Structural diversity of zinc(II) coordination polymers with octafluorobiphenyl-4,4′-dicarboxylate based on mononuclear, paddle wheel and cuboidal units. CrystEngComm 2019, 21, 2524–2533. [Google Scholar] [CrossRef]
  57. Teichert, J.; Ruck, M. Influence of Common Anions on the Coordination of Metal Cations in Polyalcohols. Eur. J. Inorg. Chem. 2019, 2019, 2267–2276. [Google Scholar] [CrossRef]
Figure 1. Redox states of anilate-type ligands.
Figure 1. Redox states of anilate-type ligands.
Crystals 12 00370 g001
Figure 2. Synthesis of compounds 13.
Figure 2. Synthesis of compounds 13.
Crystals 12 00370 g002
Figure 3. The view of molecular structure of the crystallographic unit in 1. Photo of crystals is in the inset.
Figure 3. The view of molecular structure of the crystallographic unit in 1. Photo of crystals is in the inset.
Crystals 12 00370 g003
Figure 4. The view of molecular structure of the crystallographic unit in 2. Photo of crystals is in the inset.
Figure 4. The view of molecular structure of the crystallographic unit in 2. Photo of crystals is in the inset.
Crystals 12 00370 g004
Figure 5. The view of molecular structure of the crystallographic unit in 3. Photo of crystals is in the inset.
Figure 5. The view of molecular structure of the crystallographic unit in 3. Photo of crystals is in the inset.
Crystals 12 00370 g005
Figure 6. The view of pores in crystals of 1 (left) and 2 (right) along the (010) vector. The pore volumes were evaluated using probe radius 1.2 Å and approx. grid spacing 0.7 Å. Color code: the outer side of the pores is yellow; the inner side is blue.
Figure 6. The view of pores in crystals of 1 (left) and 2 (right) along the (010) vector. The pore volumes were evaluated using probe radius 1.2 Å and approx. grid spacing 0.7 Å. Color code: the outer side of the pores is yellow; the inner side is blue.
Crystals 12 00370 g006
Figure 7. Hydrogen bonds in 1 (left) and 2 (right) crystals.
Figure 7. Hydrogen bonds in 1 (left) and 2 (right) crystals.
Crystals 12 00370 g007
Figure 8. TG curves for 1 (red line), 2 (blue line), and 3 (green line). DTA curves are shown with dashed lines.
Figure 8. TG curves for 1 (red line), 2 (blue line), and 3 (green line). DTA curves are shown with dashed lines.
Crystals 12 00370 g008
Table 1. Structure refinement details and crystal data for 13.
Table 1. Structure refinement details and crystal data for 13.
Coordiantion Polymer123
FormulaC22H38N2O8ZnC23H40N2O8ZnC20H32N2O6Zn
Formula weight523.91537.94461.84
Temperature (K)100(2)100(2)150(2)
Radiation sourcemicrofocus sealed X-ray tubesynchrotronmicrofocus sealed X-ray tube
Wavelength (Å)0.710730.745000.71073
Crystal systemMonoclinicTriclinicTriclinic
Space groupC2/cP-1P-1
a, Å19.673(3)11.2898(11)10.3884(5)
b, Å11.3396(16)11.3959(8)11.0512(6)
c, Å12.9104(18)13.0514(7)12.1460(6)
α, °9071.214(3)69.142(2)
β, °114.503(5)66.912(7)85.268(2)
γ, °9061.861(7)63.532(2)
V, A32620.7(7)1342.5(2)1161.60(10)
Z422
ρ, g/cm31.3281.3311.320
θ range, °2.126 to 30.2891.804 to 31.0642.199 to 26.768
Crystal size, mm0.360 × 0.090 × 0.0300.130 × 0.090 × 0.0500.095 × 0.030 × 0.013
μ, mm−10.9821.0821.092
Reflections
collected/unique19122/387251762/517629856/4943
No. of restraints1610
No. of parameters178325272
Rint0.0867-0.0638
GOF on F21.0531.0271.007
R1, wR2 [I > 2σ(I)]0.0524, 0.10980.0541, 0.14390.0505, 0.0937
R1, wR2 (all data)0.0791, 0.11890.0729, 0.15710.0913, 0.1047
Δρmaxρmin, e/Å30.637/−0.6720.996/−1.3150.376/−0.534
Table 2. Selected bond lengths (Å) in 13.
Table 2. Selected bond lengths (Å) in 13.
Bond 11Bond 12Bond 13
Zn(1)–O(1)2.1390(19)Zn(1)–O(1)2.133(2)Zn(1)–O(1)2.068(2)
Zn(1)–O(2)2.0481(17)Zn(1)–O(2)2.147(2)Zn(1)–O(2A)2.104(2)
Zn(1)–O(3B)2.0689(18)Zn(1)–O(3)2.0411(19)Zn(1)–O(3)2.045(2)
O(2)–C(2)1.263(3)Zn(1)–O(4A)2.080(2)Zn(1)–O(4B)2.072(2)
O(3B)–C(4B)1.265(3)Zn(1)–O(5)2.0502(19)Zn(1)–O(5)2.075(2)
C(2)–C(3)1.397(3)Zn(1)–O(6B)2.067(2)Zn(1)–O(6)2.089(2)
C(3)–C(4)1.548(3)O(3)–C(4)1.267(3)O(1)–C(1)1.266(3)
C(2)–C(4)1.396(3)O(4A)–C(6A)1.269(3)O(2A)–C(3A)1.265(3)
O(5)–C(11)1.260(3)O(3)–C(8)1.276(3)
O(6B)–C(13B)1.265(3)O(4B)–C(10B)1.262(3)
C(4)–C(5)1.401(3)C(1)–C(2)1.402(4)
C(4)–C(6A)1.545(4)C(1)–C(3A)1.546(4)
C(5)–C(6)1.405(4)C(2)–C(3)1.398(4)
C(11)–C(12)1.401(3)C(8)–C(9)1.399(4)
C(11)–C(13B)1.550(4)C(8)–C(10B)1.547(4)
C(12)–C(13)1.394(4)C(9)–C(10)1.403(4)
1 Symmetry transformations used to generate equivalent atoms: 1: (A) −x+1, y,−z+1/2; (B) −x+1,−y+1,−z+1; (C) x,−y+1, z−1/2; 2: (A) −x+1,−y+1,−z+1; (B) −x+1,−y+1,−z; 3: (A) −x,−y+1,−z+1; (B) −x, −y+1,−z+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

Trofimova, O.Y.; Maleeva, A.V.; Arsenyeva, K.V.; Klimashevskaya, A.V.; Yakushev, I.A.; Piskunov, A.V. Glycols in the Synthesis of Zinc-Anilato Coordination Polymers. Crystals 2022, 12, 370. https://doi.org/10.3390/cryst12030370

AMA Style

Trofimova OY, Maleeva AV, Arsenyeva KV, Klimashevskaya AV, Yakushev IA, Piskunov AV. Glycols in the Synthesis of Zinc-Anilato Coordination Polymers. Crystals. 2022; 12(3):370. https://doi.org/10.3390/cryst12030370

Chicago/Turabian Style

Trofimova, Olesya Y., Arina V. Maleeva, Kseniya V. Arsenyeva, Anastasiya V. Klimashevskaya, Il’ya A. Yakushev, and Alexandr V. Piskunov. 2022. "Glycols in the Synthesis of Zinc-Anilato Coordination Polymers" Crystals 12, no. 3: 370. https://doi.org/10.3390/cryst12030370

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