Next Article in Journal
Ferroelectric SnPz/In2Se3 as a Stable and Durable Non-Volatile 2D Ferroelectric Memory Material
Next Article in Special Issue
Crystal Nanoarchitectonics and Characterization of the Octahedral Iron(III)–Nitrate Complexes with Isomer Dimethylurea Ligands
Previous Article in Journal
3D Printed Microfluidic Cell for SAXS Time-Resolved Measurements of the Structure of Protein Crystallization Solutions
Previous Article in Special Issue
Electrochemical Properties and Perspectives of Nickel(II) and Cobalt(II) Coordination Polymers-Aspects and an Application in Electrocatalytic Oxidation of Methanol
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Two-Dimensional Lattices with Lanthanoids, Anilato Ligands and Formamide

by
Samia Benmansour
*,
Antonio Hernández-Paredes
,
Kilian Defez-Aznar
and
Carlos J. Gómez-García
*
Departamento de Química Inorgánica, Universidad de Valencia, C/Dr. Moliner 50, 46100 Burjasot, Valencia, Spain
*
Authors to whom correspondence should be addressed.
Crystals 2023, 13(6), 939; https://doi.org/10.3390/cryst13060939
Submission received: 28 April 2023 / Revised: 23 May 2023 / Accepted: 7 June 2023 / Published: 11 June 2023
(This article belongs to the Special Issue Coordination Complexes: Synthesis, Characterization and Application)

Abstract

:
Here, we illustrate the use of formamide (fma) and anilato-type ligands to build two-dimensional lattices with lanthanoids. Thus, we describe the synthesis and crystal structure of four lattices formulated as [Ln2(C6O4X2)3(fma)6]·6fma with Ln/X = La/Cl (1), La/Br (2), Eu/Cl (3), and Eu/Br (4), where C6O4X22− = dianion of 3,6-disubstituted-2,5-dihydroxy-1,4-benzoquinone with X = Cl (chloranilato) and X = Br (bromanilato). Single crystal X-ray analysis shows that the four compounds crystallize in the triclinic P-1 space group and present two-dimensional, very distorted hexagonal lattices with the lanthanoids ions in the vertex coordinated by three anilato ligands forming the sides of the distorted hexagons that appear as rectangles. The rectangles are disposed parallel to their long sides in a brick wall fashion. The nona-coordination of the lanthanoids is completed by three formamide molecules. These layered compounds include three additional formamide molecules per lanthanoid atom, located in the interlayer space inside the channels formed by the eclipsed packing of the layers. We discuss the differences observed among these compounds due to the change of the lanthanoid ion (La and Eu) and of the substituent group X in the anilato ligand (Cl and Br).

1. Introduction

According to IUPAC [1], MOFs can be defined as coordination networks formed by organic ligands containing potential voids. These porous coordination networks have experienced a huge interest in the 21st century, resulting in the synthesis of several thousands of MOFs with large porosities and surface areas, including examples with flexible skeletons and chemically functionalized cavities [2]. Of course, the control and modulation of the pore sizes and shapes, in order to modulate the properties of the MOFs, constitutes one of the most challenging tasks in the synthesis of these materials [3,4].
Although most MOFs contain transition metal atoms, there is an increasing interest in the synthesis of lanthanoid-based metal–organic frameworks (Ln-MOFs). This interest is based on the many properties and potential applications that they show. Thus, besides the well-known applications of many MOFs, such as gas storage and separation [5], catalysis [6], magnetic, optical, and chemical sensing [7,8,9], water absorption [10], etc., Ln-MOFs present additional properties, such as magnetism, including single-molecule and single-ion magnets (SMM and SIM), that makes them potential candidates for other applications in spintronics and quantum information processing [11] or as luminescent-based chemical sensors [12].
Besides three-dimensional (3D) MOFs, the synthesis of two-dimensional (2D) MOFs has gained a lot of attention in the last years, most likely due to the discovery of graphene and other 2D layered materials. These efforts have yielded several 2D MOFs with interesting properties and applications [13,14].
Many different ligands have been used to construct these MOFs. Some of them are from the family of 3,6-disubstituted-2,5-dihydroxy-1,4-benzoquinone dianions = C6O4X22−, known as anilato ligands (Scheme 1a). These ligands present interesting advantages as they show different coordination modes such as: bidentate terminal (1k2O,O′, Scheme 1b), bis-bidentate (1k2O,O′;2k2O″,O‴, Scheme 1c), monodentate (1kO), monodentate-bidentate (1kO;2k2O′,O″), and even more complex modes (1k2O,O′;2k2O″,O‴;3kO″) [15]. The bidentate and bis-bidentate coordination modes are the most common ones and, as observed for the topologically equivalent oxalate ligand, result in the same type of structures from mononuclear tris(anilato)metalate complexes [16,17] to polymeric 1D, 2D, and 3D lattices, although with larger cavities and channels [15,18,19,20,21,22].
Anilato-based MOFs with lanthanoids constitute a growing family since the very first reports of Raymond et al. [23], and, especially, Robson and Abrahams et al. [24,25]. A recent review by some of us [22] showed that there are more than 140 known compounds prepared with anilato ligands (C6O4X22− with X = H, F, Cl, Br, Cl/CN, CH3, NO2, and t-Bu) with all the lanthanoids ions (except Pm) and different solvents acting as co-ligands, such as H2O [24,25], dimethyl sulfoxide [26,27], dimethylformamide [28,29], dimethylacetamide [30], ethanol [23], ethylene glycol [31], etc. Surprisingly, formamide has been used only in one case [32] among the more than 140 anilato–lanthanoid compounds reported to date [22].
The several families of compounds prepared with different solvents, lanthanoids, and anilato ligands [22] have clearly demonstrated that the final topology and structure depend on several factors: (i) the size of the LnIII ion, (ii) the X groups of the anilato ligands, (iii) the shape and size of the coordinated solvent molecules, (iv) the synthetic method, and even (v) post synthetic solvent exchange treatment [33].
As there is only one reported example with formamide as solvent: [Er2(C6O4Cl2)3(fma)6]·4fma·2H2O [32], we have explored the possibility of preparing other lattices with this solvent and different lanthanoid metal ions using chloranilato and bromanilato (X = Cl and Br, Scheme 1). These attempts have led to the synthesis of four isostructural compounds formulated as [Ln2(C6O4X2)3(fma)6]·6fma with Ln/X = La/Cl (1), La/Br (2), Eu/Cl (3), and Eu/Br (4). Here, we show the synthesis and structure of the four compounds and a detailed study of the differences observed in these four compounds caused by the change in X and in the lanthanoid ion.

2. Materials and Methods

Chloranilic acid, bromanilic acid, La(NO3)3·6H2O, Eu(NO3)3·6H2O, and fma are all commercial and were used as received without any further purification. All the compounds were obtained as single crystals using a layering method described below.
Synthesis of [La2(C6O4Cl2)3(fma)6]·6fma (1) Single crystals of compound 1 were obtained by carefully layering, at room temperature, a solution of chloranilic acid, H2C6O4Cl2 (6.2 mg, 0.03 mmol) in 6.7 mL of methanol on top of a solution of La(NO3)3.6H2O (8.7 mg, 0.02 mmol) in 4.5 mL of fma. The tube (300 mm length and 5 mm diameter) was sealed and allowed to stand for one week to obtain purple prismatic single crystals suitable for X-ray diffraction that were freshly picked and covered with paratone oil (to avoid solvent loss) to be characterized by single crystal X-ray diffraction.
Synthesis of [La2(C6O4Br2)3(fma)6]·6fma (2) Purple prismatic single crystals of compound 2 were obtained in the same way as 1, but using bromanilic acid, H2C6O4Br2 (6.0 mg, 0.02 mmol) in 4.5 mL of methanol, instead of chloranilic acid.
Synthesis of [Eu2(C6O4Cl2)3(fma)6]·6fma (3) Purple prismatic single crystals of compound 3 were obtained in the same way as 1, but using Eu(NO3)3.6H2O (8.9 mg, 0.02 mmol) instead of La(NO3)3·6H2O.
Synthesis of [Eu2(C6O4Cl2)3(fma)6]·6fma (4) Purple prismatic single crystals of compound 4 were obtained in the same way as 2, but using Eu(NO3)3.6H2O (8.9 mg, 0.02 mmol) instead of La(NO3)3·6H2O.
IR spectroscopy. IR spectra were performed on KBr pellets with a Bruker Equinox 55 spectrometer in the wavelength range 400–4000 cm−1.
X-ray crystallography. Single crystals of all compounds were mounted on glass fibres with a viscous hydrocarbon oil to coat the crystals. The crystals were immediately transferred to a cold N2 stream for data collection. X-ray data were collected on a Supernova diffractometer equipped with a graphite monochromated Enhance (Mo) X-ray Source (λ = 0.71073 Å) at 120 K. The program CrysAlisPro, Oxford Diffraction Ltd., was used for unit cell determinations and data reduction [34]. Empirical absorption correction was performed using spherical harmonics, implemented in the SCALE3 ABSPACK scaling algorithm. All compounds crystallize in the triclinic P-1 space group. Crystal structures were solved by direct methods with the SIR92 program [35] and refined against all F2 values with the SHELXL-2014 program [36], using the WinGX2014.1 graphical user interface [37]. Non-hydrogen atoms were refined anisotropically, and hydrogen atoms were assigned fixed isotropic displacement parameters. Hydrogen atom positions were calculated geometrically and refined using the riding model. For the four compounds, hydrogen atoms of the ligands were set in calculated positions and refined as riding atoms, whereas the H atoms of crystallized formamide molecules, in some cases, could not be found nor calculated. For compound 2, the use of some bond lengths restraints, applied on atoms belonging to formamide solvent molecules, has been reasonably imposed like DFIX and ISOR. In compound 3, the use of some bond lengths restraints, applied on atoms belonging to dynamic moieties, has been reasonably imposed like RIGU and ISOR. Regarding compound 4, the use of some bond lengths restraints, applied on atoms belonging to formamide solvent, has been reasonably imposed like ISOR and DFIX.
Complete tables with bond distances for compounds 14 are provided in the Supplementary Materials. A summary of the data collection and structure refinement for compounds 14 is given in Table 1 and Table 2. Crystallographic data for the structures reported in this paper have been deposited in the Cambridge Crystallographic Data Centre with CCDC numbers 2259696–2259699. These data can be downloaded via www.ccdc.cam.ac.uk/[email protected] (accessed on 1 April 2023).

3. Results

3.1. Synthesis and IR Spectra

The synthesis of compounds 14 was performed following the same layering method but using different Ln(NO3)3·nH2O with chloranilic and bromanilic acids, and fma and methanol as solvents. As expected, since amides have a larger affinity towards lanthanoids than methanol [38], the resulting structures contain LnIII ions coordinated to formamide rather than to methanol. The IR spectra of the four compounds are very similar, as expected, given their isostructurality and close composition. They show the expected bands associated with the anilato ligands and fma molecules. The main differences are observed in the C-X bands (X = Cl and Br) of the anilato ligands (see supporting information).

3.2. Crystal Structure of [Ln2(C6O4X2)3(fma)6]·6fma with Ln/X = La/Cl(1), La/Br(2), Eu/Cl(3), and Eu/Br(4)

Compounds 14 are isostructural and, therefore, only the structure of compound 3 will be described in detail with a comparative study for the other compounds. Compounds 14 crystallize in the triclinic P-1 space group. The asymmetric unit (Figure 1a) contains one LnIII ion, three half anilato ligands, three coordinated fma molecules, and three crystallization fma molecules. The general formula is, then, [Ln2(C6O4X2)3(fma)6]·6fma with Ln/X = La/Cl (1), La/Br (2), Eu/Cl (3), and Eu/Br (4).
Each metal centre is coordinated by three chelating anilato ligands and completes its nona-coordination with three fma molecules. These three anilato ligands connect each LnIII with three other metal centres (Figure 1b). The LnIII ion is surrounded by nine O atoms: O2, O6, O12, O16, O22, and O26, from the three chelating anilato ligands, and O1D, O11D, and O21D, from the three coordinated fma molecules (Figure 1c). Continuous SHAPE analysis of the coordination environment shows that the coordination geometry of the metal centres is a slightly distorted caped square antiprism (CSAPR-9, Table 3) [39].
A close look at the coordination environment of the LnIII ions (Figure 1c) shows that the fma O atoms occupy the upper square face of the CSAPR. We call this a 030 location, (where the two first digits refer to the number of solvent molecules in the lower and upper square faces, respectively, and the third one to the capping position). As we will see below, the exact position of the O atoms of the fma molecules and of the anilato ligands determines the spatial orientation of the bridging anilato ligands and, thus, the distortions of the hexagonal cavities and the final structure.
The connectivity of the metal ions through the anilato ligands gives rise to a two-dimensional structure with a (6,3)-gon topology where each metal atom is connected to three other metal atoms, forming very distorted, almost rectangular, six-membered rings, disposed in flat layers and formulated as [Ln2(C6O4X2)3(fma)6] (Figure 2). The crystallization fma molecules are located in between these layers as well as inside the layers in the cavities. The coordinated fma molecules are located orthogonal to the layers, pointing towards the interlayer space (Figure 2).
The cavities are arranged in parallel to their long axis with a brick wall disposition (Figure 3a). In each cavity, four of the anilato ligands are disposed perpendicular (edge on) and two perpendicular (face on) to the cavity plane (Figure 3a). The layers are packed parallel to the bc plane in an eclipsed way, giving rise to channels running along the a direction (Figure 3b).

4. Discussion

The structure of compounds 14 is similar to the c-type structure reported by Robson et al. [23,25], with water as coordinated solvent, and has the same (6,3)-gon topology found in other Ln-containing anilato-based lattices [22].
If we compare in detail the structures of compounds 14, we can observe that the average Ln-Ofma bond distances are shorter than the Ln-Oanilato ones in all cases (Table 4). This difference can be attributed to the rigidity of the five-membered chelate rings formed with the anilato ligands. As expected, the La-O bond distances (1 and 2) are longer than the Eu-O ones (3 and 4), since LaIII is larger than EuIII. If we compare the two compounds with the same LnIII ion, we can see that the Ln-Oanilato bond distances for the chloranilato compounds are slightly shorter than the corresponding ones with bromanilato. Although the difference is small, this difference has to be attributed to the larger steric hindrance of the Br atoms, since the electronic effects should lead to the opposite trend (the O atoms of chloranilato have less electron density and, therefore, should form longer Ln-O bonds).
The distortions of the hexagonal cavities in 3 can be quantified by the three Ln-Ln distances along the diagonals of the distorted hexagonal cavity (20.74, 17.78, and 11.36 Å), as well as by the three Ln-Ln-Ln angles (87.30, 109.02, and 159.18°). These values in the other compounds are similar, although, as expected, the Ln-Ln distances are longer in the LaIII compounds (1 and 2) than in the EuIII ones (3 and 4, Table 5). On the other side, if we compare the chloranilato and bromanilato derivatives for the same lanthanoid, we can see that in the bromanilato derivatives (2 and 4), the Ln-Ln distances are slightly longer than in the corresponding chloranilato ones (1 and 3), as a consequence of the larger steric hindrance of Br. In contrast, the Ln-Ln-Ln angles do not show a general trend since they are correlated and the decrease of one of them leads to an increase of the others.
As can be observed in Table 3, although the coordination geometry of the LnIII ions is caped square antiprism (CSAPR-9) in all cases, the two compounds with LaIII show higher SHAPE parameter values (0.478 in 1 and 0.487 in 2) than the EuIII derivatives (0.318 in 3 and 0.348 in 4). The larger distortions in the LaIII derivatives are due to the larger size of LaIII compared to EuIII and has already been observed in other anilato-based compounds with other solvent molecules [26]. Furthermore, for both lanthanoids, the compounds with bromanilato (2 and 4) show larger distortions from the ideal geometry than the corresponding chloranilato ones (1 and 3). This fact reflects the larger steric hindrance of the Br atoms in bromanilato compared to Cl in chloranilato.
Although LaIII and EuIII have different sizes, they present the same coordination number (nine) and geometry (CSAPR-9). This fact can be explained by the small size and steric hindrance of the fma molecule that allows the coordination of three molecules, even for the smaller EuIII ion. Note also that although the larger size of LaIII might lead to a coordination number of ten, this high coordination number has only been observed once with anilato ligands and LaIII, and in that case, the additional ligand was H2O [22,25].

5. Conclusions

We have synthesized and structurally characterized four isostructural anilato-based compounds with LaIII and EuIII and formamide as solvent. These compounds, formulated as [Ln2(C6O4X2)3(fma)6]·6fma with Ln/X = La/Cl (1), La/Br (2), Eu/Cl (3), and Eu/Br (4), with C6O4X22− = chloranilato (X = Cl) and bromanilato (X = Br), crystallize in the triclinic P-1 space group. The LnIII centres are surrounded by nine oxygen atoms from three chelating anilato ligands and three coordinated formamide molecules. The coordination geometry is slightly distorted caped square antiprism with a 030 disposition of the three fma solvent molecules. The spatial orientation of the anilato ligands around the metal ions leads to a two-dimensional structure with a (6,3)-gon topology with very distorted hexagonal cavities that appear as rectangles showing a brick wall disposition in the layers. There are formamide crystallization molecules located in between the layers and in the rectangular channel formed by the eclipsed packing of the layers.
A detailed comparative analysis in compounds 14 of the influence of the size of the lanthanoid and of the X group of the anilato ligand in the structural parameters, shows that: (i) the LaIII compounds present a higher distortion of the coordination geometry than the EuIII ones, (ii) the bromanilato compounds present a higher distortion than the corresponding chloranilato ones, (iii) the LnIII-O bond distances are larger for the LaIII compounds than for the EuIII ones, (iv) the LnIII-O bond distances are larger for the bromanilato derivatives than for the chloranilato ones, and (v) the distortions of the rectangular cavities are larger for the LaIII compounds and for the bromanilato derivatives. All these differences can be explained by the larger size of LaIII compared to EuIII and by the larger steric hindrance of Br.
The results reported here open the gate to the synthesis of the corresponding compounds with other lanthanoids to confirm the influence of the lanthanoid size and to check for possible changes in the coordination number and in the structure as the LnIII size decreases. On the other hand, an optical study is ongoing for the four compounds. We are also preparing the DyIII and ErIII derivatives with both anilato ligands in order to obtain single-ion magnet behaviour, as has already been observed in other anilato-based DyIII and ErIII compounds with different solvents [22]. Work is in progress.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/cryst13060939/s1, Tables S1–S4: main bond distances (in Å) in compounds 14, respectively. Table S5. Main IR bands and their assignments in compounds 14. Figures S1–S4: IR spectra of compounds 14, respectively.

Author Contributions

Conceptualization, S.B. and C.J.G.-G.; data curation, S.B., C.J.G.-G., A.H.-P. and K.D.-A.; formal analysis, S.B. and C.J.G.-G.; funding acquisition, C.J.G.-G.; investigation, S.B., C.J.G.-G. and A.H.-P.; methodology, S.B., C.J.G.-G., A.H.-P. and K.D.-A.; project administration, C.J.G.-G.; supervision, S.B. and C.J.G.-G.; writing—original draft, C.J.G.-G.; writing—review and editing, S.B., C.J.G.-G. and A.H.-P. All authors have read and agreed to the published version of the manuscript.

Funding

This study forms part of the Advanced Materials program and was supported by the Spanish MCIN with funding from European Union NextGeneration EU (PRTR- C17.I1) and the Generalitat Valenciana (project MFA-2022-057). We also thank the project PID2021-125907NB-I00, financed by MCIN/AEI/10.13039/50110 0 011033/FEDER, UE, for financial support.

Data Availability Statement

The data presented in this study are available on request from the corresponding author.

Conflicts of Interest

The authors declare no conflict of interest. The funders had no role in the design of the study; in the collection, analyses, or interpretation of data; in the writing of the manuscript, or in the decision to publish the results.

References

  1. Batten, S.R.; Champness, N.R.; Chen, X.; Garcia-Martinez, J.; Kitagawa, S.; Öhrström, L.; O’Keeffe, M.; Suh, M.P.; Reedijk, J. Terminology of metal–organic Frameworks and Coordination Polymers (IUPAC Recommendations 2013). Pure Appl. Chem. 2013, 85, 1715–1724. [Google Scholar] [CrossRef] [Green Version]
  2. Zhang, X.; Chen, Z.; Liu, X.; Hanna, S.L.; Wang, X.; Taheri-Ledari, R.; Maleki, A.; Li, P.; Farha, O.K. A Historical Overview of the Activation and Porosity of metal–organic Frameworks. Chem. Soc. Rev. 2020, 49, 7406–7427. [Google Scholar] [CrossRef]
  3. Zhou, H.C.; Long, J.R.; Yaghi, O.M. Introduction to Metal-Organic Frameworks. Chem. Rev. 2012, 112, 673–674. [Google Scholar] [CrossRef]
  4. Furukawa, H.; Cordova, K.E.; O’Keeffe, M.; Yaghi, O.M. The Chemistry and Applications of Metal-Organic Frameworks. Science 2013, 341, 1230444. [Google Scholar] [CrossRef] [Green Version]
  5. Zhao, X.; Wang, Y.; Li, D.; Bu, X.; Feng, P. Metal–Organic Frameworks for Separation. Adv. Mater. 2018, 30, 1705189. [Google Scholar] [CrossRef] [PubMed]
  6. Shu-Na, Z.; Wang, G.; Poelman, D.; Van Der Voort, P. Metal Organic Frameworks Based Materials for Heterogeneous Photocatalysis. Molecules 2018, 23, 2947. [Google Scholar]
  7. Li, H.; Zhao, S.; Zang, S.; Li, J. Functional metal–organic Frameworks as Effective Sensors of Gases and Volatile Compounds. Chem. Soc. Rev. 2020, 49, 6364–6401. [Google Scholar] [CrossRef] [PubMed]
  8. Kreno, L.E.; Leong, K.; Farha, O.K.; Allendorf, M.; Van Duyne, R.P.; Hupp, J.T. Metal-Organic Framework Materials as Chemical Sensors. Chem. Rev. 2012, 112, 1105–1125. [Google Scholar] [CrossRef]
  9. Campbell, M.G.; Dinca, M. Metal-Organic Frameworks as Active Materials in Electronic Sensor Devices. Sensors 2017, 17, 1108. [Google Scholar] [CrossRef] [Green Version]
  10. Canivet, J.; Fateeva, A.; Guo, Y.; Coasne, B.; Farrusseng, D. Water Adsorption in MOFs: Fundamentals and Applications. Chem. Soc. Rev. 2014, 43, 5594–5617. [Google Scholar] [CrossRef] [Green Version]
  11. Marin, R.; Brunet, G.; Murugesu, M. Shining New Light on Multifunctional Lanthanide Single-Molecule Magnets. Angew. Chem. Int. Ed. 2021, 60, 1728–1746. [Google Scholar] [CrossRef] [PubMed]
  12. Rocha, J.; Carlos, L.D.; Paz, F.A.A.; Ananias, D. Luminescent Multifunctional Lanthanides-Based metal–organic Frameworks. Chem. Soc. Rev. 2011, 40, 926–940. [Google Scholar] [CrossRef] [PubMed]
  13. Chakraborty, G.; Park, I.; Medishetty, R.; Vittal, J.J. Two-Dimensional Metal-Organic Framework Materials: Synthesis, Structures, Properties and Applications. Chem. Rev. 2021, 121, 3751–3891. [Google Scholar] [CrossRef] [PubMed]
  14. 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]
  15. Benmansour, S.; Vallés-García, C.; Gómez-Claramunt, P.; Mínguez Espallargas, G.; Gómez-García, C.J. 2D and 3D Anilato-Based Heterometallic M(I)M(III) Lattices: The Missing Link. Inorg. Chem. 2015, 54, 5410–5418. [Google Scholar] [CrossRef]
  16. Atzori, M.; Artizzu, F.; Sessini, E.; Marchio, L.; Loche, D.; Serpe, A.; Deplano, P.; Concas, G.; Pop, F.; Avarvari, N.; et al. Halogen-Bonding in a New Family of Tris(Haloanilato)Metallate(III) Magnetic Molecular Building Blocks. Dalton Trans. 2014, 43, 7006–7019. [Google Scholar] [CrossRef]
  17. Benmansour, S.; Gómez-Claramunt, P.; Vallés-García, C.; Mínguez Espallargas, G.; Gómez García, C.J. Key Role of the Cation in the Crystallization of Chiral Tris(Anilato)Metalate Magnetic Anions. Cryst. Growth Des. 2016, 16, 518–526. [Google Scholar] [CrossRef]
  18. 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]
  19. Abrahams, B.F.; Grannas, M.J.; Hudson, T.A.; Hughes, S.A.; Pranoto, N.H.; Robson, R. Synthesis, Structure and Host-Guest Properties of (Et4N)2[SnIVCaII(Chloranilate)4], a New Type of Robust Microporous Coordination Polymer with a 2D Square Grid Structure. Dalton Trans. 2011, 40, 12242–12247. [Google Scholar] [CrossRef]
  20. Benmansour, S.; Gómez-García, C.J. A Heterobimetallic Anionic 3,6-Connected 2D Coordination Polymer Based on Nitranilate as Ligand. Polymers 2016, 8, 89. [Google Scholar] [CrossRef] [Green Version]
  21. 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]
  22. Benmansour, S.; Gómez-García, C.J. Lanthanoid-Anilato Complexes and Lattices. Magnetochemistry 2020, 6, 71. [Google Scholar] [CrossRef]
  23. Riley, P.E.; Haddad, S.F.; Raymond, K.N. Preparation of Praseodymium(III) Chloranilate and the Crystal Structures of Pr2(C6Cl2O4)3·8C2H5OH and Na3[C6H2O(OH)(SO3)2]·H2O. Inorg. Chem. 1983, 22, 3090–3096. [Google Scholar] [CrossRef]
  24. Abrahams, B.F.; Coleiro, J.; Hoskins, B.F.; Robson, R. Gas Hydrate-like Pentagonal Dodecahedral M2(H2O)18 cages (M = Lanthanide or Y) in 2,5-Dihydroxybenzoquinone-Derived Coordination Polymers. Chem. Commun. 1996, 5, 603–604. [Google Scholar] [CrossRef]
  25. Abrahams, B.F.; Coleiro, J.; Ha, K.; Hoskins, B.F.; Orchard, S.D.; Robson, R. Dihydroxybenzoquinone and Chloranilic Acid Derivatives of Rare Earth Metals. J. Chem. Soc. Dalton Trans. 2002, 8, 1586–1594. [Google Scholar] [CrossRef]
  26. Benmansour, S.; Gómez-García, C.J.; Hernández-Paredes, A. The Complete Series of Lanthanoid-Chloranilato Lattices with Dimethylsulfoxide: Role of the Lanthanoid Size on the Coordination Number and Crystal Structure. Crystals 2022, 12, 261. [Google Scholar] [CrossRef]
  27. Benmansour, S.; Hernández-Paredes, A.; Gómez-García, C.J. Effect of the Lanthanoid-Size on the Structure of a Series of Lanthanoid-Anilato 2-D Lattices. J. Coord. Chem. 2018, 71, 845–863. [Google Scholar] [CrossRef]
  28. Gómez-Claramunt, P.; Benmansour, S.; Hernández-Paredes, A.; Cerezo-Navarrete, C.; Rodríguez-Fernández, C.; Canet-Ferrer, J.; Cantarero, A.; Gómez-García, C.J. Tuning the Structure and Properties of Lanthanoid Coordination Polymers with an Asymmetric Anilato Ligand. Magnetochemistry 2018, 4, 6. [Google Scholar] [CrossRef] [Green Version]
  29. Benmansour, S.; Hernández-Paredes, A.; Gómez-García, C.J. Two-Dimensional Magnetic Coordination Polymers Formed by Lanthanoids and Chlorocyananilato. Magnetochemistry 2018, 4, 58. [Google Scholar] [CrossRef] [Green Version]
  30. 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 Ligands. CrystEngComm 2020, 22, 4675–4679. [Google Scholar] [CrossRef]
  31. Benmansour, S.; Pintado-Zaldo, C.; Martínez-Ponce, J.; Hernández-Paredes, A.; Valero-Martínez, A.; Gómez-Benmansour, M.; Gómez-García, C.J. The Versatility of Ethylene Glycol to Tune the Dimensionality and Magnetic Properties in DyIII-Anilato-Based Single-Ion Magnets. Cryst. Growth Des. 2023, 23, 1269–1280. [Google Scholar] [CrossRef]
  32. Benmansour, S.; Pérez-Herráez, I.; Cerezo-Navarrete, C.; López-Martínez, G.; Martínez Hernandez, C.; Gómez-García, C.J. Solvent-Modulation of the Structure and Dimensionality in Lanthanoid-Anilato Coordination Polymers. Dalton Trans. 2018, 47, 6729–6741. [Google Scholar] [CrossRef] [PubMed]
  33. Benmansour, S.; Hernández-Paredes, A.; Mondal, A.; López Martínez, G.; Canet-Ferrer, J.; Konar, S.; Gómez-García, C.J. Slow Relaxation of the Magnetization, Reversible Solvent Exchange and Luminescence in 2D Anilato-Based Frameworks. Chem. Commun. 2020, 56, 9862–9865. [Google Scholar] [CrossRef] [PubMed]
  34. Oxford Diffraction CrysalisPro. 2004. 171.33.55. Available online: https://www.rigaku.com/products/crystallography/crysalis (accessed on 27 April 2023).
  35. Altomare, A.; Burla, M.C.; Camalli, M.; Cascarano, G.L.; Giacovazzo, C.; Guagliardi, A.; Moliterni, A.G.G.; Polidori, G.; Spagna, R. SIR97: A New Tool for Crystal Structure Determination and Refinement. J. Appl. Cryst. 1999, 32, 115–119. [Google Scholar] [CrossRef]
  36. Sheldrick, G.M. Crystal Structure Refinement with SHELXL. Acta Cryst. C 2015, 71, 3–8. [Google Scholar] [CrossRef] [Green Version]
  37. Farrugia, L.J. WinGX and ORTEP for Windows: An Update. J. Appl. Cryst. 2012, 45, 849–854. [Google Scholar] [CrossRef]
  38. Álvarez, S. Coordinating Ability of Anions, Solvents, Amino Acids, and Gases Towards Alkaline and Alkaline-Earth Elements, Transition Metals, and Lanthanides. Chem. Eur. J. 2020, 26, 4350–4377. [Google Scholar] [CrossRef]
  39. Álvarez, S. Distortion Pathways of Transition Metal Coordination Polyhedra Induced by Chelating Topology. Chem. Rev. 2015, 115, 13447–13483. [Google Scholar] [CrossRef] [Green Version]
Scheme 1. (a) Anilato-type ligands and main coordination modes: (b) terminal bidentate and (c) bridging bis-bidentate.
Scheme 1. (a) Anilato-type ligands and main coordination modes: (b) terminal bidentate and (c) bridging bis-bidentate.
Crystals 13 00939 sch001
Figure 1. (a) ORTEP plot of the asymmetric unit of compound 3 showing the labelling scheme. Crystallization fma molecules are omitted for clarity. Ellipsoids are drawn at 80% probability. (b) Coordination environment around the Eu centre in compound 3. Colour code: Eu = pink, C = grey, O = red, N = blue, Cl = green, and H = white. (c) Caped square antiprism coordination geometry around the Eu ion in compound 3. The O atoms of the fma molecules are drawn in orange.
Figure 1. (a) ORTEP plot of the asymmetric unit of compound 3 showing the labelling scheme. Crystallization fma molecules are omitted for clarity. Ellipsoids are drawn at 80% probability. (b) Coordination environment around the Eu centre in compound 3. Colour code: Eu = pink, C = grey, O = red, N = blue, Cl = green, and H = white. (c) Caped square antiprism coordination geometry around the Eu ion in compound 3. The O atoms of the fma molecules are drawn in orange.
Crystals 13 00939 g001
Figure 2. View of three consecutive layers in compound 3 showing the coordinated fma molecules pointing towards the interlayer space and the crystallization fma molecules located in between the layers. Colour code: Eu = pink, C = grey, O = red, N = blue, Cl = green, and H = white.
Figure 2. View of three consecutive layers in compound 3 showing the coordinated fma molecules pointing towards the interlayer space and the crystallization fma molecules located in between the layers. Colour code: Eu = pink, C = grey, O = red, N = blue, Cl = green, and H = white.
Crystals 13 00939 g002
Figure 3. (a) View, down the a direction, of one brick wall-like layer parallel to the bc plane in compound 3 showing the crystallization fma molecules located above and below (light blue), and inside (purple) the layers. (b) Perspective view of one rectangular channel along the a direction in compound 3 showing the fma molecules inside the channels. Colour code in (b): Eu = pink, C = grey, O = red, N = blue, Cl = green, and H = white.
Figure 3. (a) View, down the a direction, of one brick wall-like layer parallel to the bc plane in compound 3 showing the crystallization fma molecules located above and below (light blue), and inside (purple) the layers. (b) Perspective view of one rectangular channel along the a direction in compound 3 showing the fma molecules inside the channels. Colour code in (b): Eu = pink, C = grey, O = red, N = blue, Cl = green, and H = white.
Crystals 13 00939 g003
Table 1. Crystal data and structure refinement parameters for compounds [La2(C6O4X2)3(fma)6]·6fma with X = Cl(1) and Br(2).
Table 1. Crystal data and structure refinement parameters for compounds [La2(C6O4X2)3(fma)6]·6fma with X = Cl(1) and Br(2).
Compound[La2(C6O4Cl2)3(fma)6]·6fma[La2(C6O4Br2)3(fma)6]·6fma
Ref12
CCDC22596962259697
Empirical formulaC15H5Cl3LaN6O12C15H8Br3LaN6O12
Formula weight706.51842.91
Crystal systemtriclinictriclinic
Space groupP-1P-1
a (Å)8.9314(10)8.9686(9)
b (Å)10.5972(9)10.7072(12)
c (Å)14.5498(13)14.5975(11)
α (°)75.639(7)75.178(8)
β (°)84.309(8)83.957(7)
γ (°)76.755(8)76.664(9)
Volume (Å3)1297.3(2)1317.0(2)
Z22
Density (calculated) (g/cm3)1.8092.126
Absorption coefficient (mm−1)2.0226.239
F(000)682796
Crystal size (mm3)0.07 × 0.04 × 0.020.07 × 0.04 × 0.02
2θ range for data (°)3.372–25.0483.367–25.036
Reflections collected152268761
Data45824651
Restraints051
Parameters343352
Goodness-of-fit on F21.0621.052
R1 [I > 2s(I)]0.04880.0452
wR2 (all data)0.12280.0954
Largest diff. peak/hole/e Å−31.126 and −0.8971.278 and −0.954
Table 2. Crystal data and structure refinement parameters for compounds [Eu2(C6O4X2)3(fma)6]·6fma with X = Cl (3) and Br (4).
Table 2. Crystal data and structure refinement parameters for compounds [Eu2(C6O4X2)3(fma)6]·6fma with X = Cl (3) and Br (4).
Compound[Eu2(C6O4Cl2)3(fma)6]·6fma[Eu2(C6O4Br2)3(fma)6]·6fma
Ref34
CCDC22596982259699
Empirical formulaC15H5Cl3EuN6O12C15H8Br3EuN6O12
Formula weight729.64861.00
Crystal systemtriclinictriclinic
Space groupP-1P-1
a (Å)8.8487(4)8.8978(6)
b (Å)10.5840(5)10.6669(7)
c (Å)14.2341(6)14.3447(10)
α (°)75.534(4)75.168(6)
β (°)84.718(4)84.539(5)
γ (°)76.350(4)76.145(6)
Volume (Å3)1253.59(10)1276.99(16)
Z22
Density (calculated) (g/cm3)1.9332.239
Absorption coefficient (mm−1)2.8917.218
F(000)714818
Crystal size (mm3)0.11 × 0.08 × 0.040.09 × 0.05 × 0.02
2θ range for data (°)3.309–25.0486.75–50.1
Reflections collected44328637
Data40084496
Restraints4274
Parameters343334
Goodness-of-fit on F21.0471.032
R1 [I > 2s(I)]0.02960.0496
wR2 (all data)0.06620.1223
Largest diff. peak/hole/e Å−31.495 and −0.9562.15 and −2.55
Table 3. Continuous SHAPE measurement (CShM) values of the 13 possible coordination geometries for the LnIII ion with coordination number nine in compounds 14. The minimum values are indicated in bold.
Table 3. Continuous SHAPE measurement (CShM) values of the 13 possible coordination geometries for the LnIII ion with coordination number nine in compounds 14. The minimum values are indicated in bold.
GeometrySymmetry1 (La/Cl)2 (La/Br)3 (Eu/Cl)4 (Eu/Br)
EP-9D9h35.52635.41935.98835.960
OPY-9C8v22.64522.74222.28622.320
HBPY-9D7h19.61719.30919.75319.516
JTC-9C3v14.62914.64215.02114.981
JCCU-9C4v10.0759.9469.9149.827
CCU-9C4v8.9828.8648.9328.842
JCSAPR-9C4v1.3891.3831.1201.146
CSAPR-9C4v0.4780.4870.3180.348
JTCTPR-9D3h2.8662.9862.6442.825
TCTPR-9D3h1.5921.6181.3691.490
JTDIC-9C3v12.09112.15112.27312.234
HH-9C2v11.16211.11111.64111.504
MFF-9Cs0.9690.9750.8820.870
EP-9 = Enneagon; OPY-9 = Octagonal pyramid; HBPY-9 = Heptagonal bipyramid; JTC-9 = Triangular cupola (J3 = trivacant cuboctahedron); JCCU-9 = Capped cube (Elongated square pyramid, J8); CCU-9 = Capped cube; JCSAPR-9 = Capped square antiprism (Gyroelongated square pyramid J10); CSAPR-9 = Capped square antiprism; JTCTPR-9 = Tricapped trigonal prism (J51); TCTPR-9 = Tricapped trigonal prism; JTDIC-9 = Tridiminished icosahedron (J63); HH-9 = Hula hoop; MFF-9 = Muffin.
Table 4. Ln-O bond distances (in Å) in compounds 14.
Table 4. Ln-O bond distances (in Å) in compounds 14.
Atoms1 (La/Cl)2 (La/Br)3 (Eu/Cl)4 (Eu/Br)
Ln-O22.557(4)2.622(5)2.454(3)2.453(5)
Ln-O62.614(4)2.566(4)2.465(3)2.453(5)
Ln-O122.548(4)2.556(5)2.455(3)2.465(5)
Ln-O162.552(4)2.557(5)2.548(3)2.548(5)
Ln-O222.526(4)2.544(4)2.444(3)2.477(5)
Ln-O262.544(4)2.539(4)2.459(3)2.483(5)
Ln-O1D2.503(4)2.520(4)2.377(3)2.382(5)
Ln-O11D2.468(4)2.508(5)2.418(3)2.413(5)
Ln-O21D2.508(4)2.483(4)2.400(3)2.429(5)
Ln-Oanilato 12.5572.5642.4712.480
Ln-Ofma 22.4932.5042.3982.408
1 Average Ln-O bond distance of the anilato oxygen atoms: O2, O6, O12, O16, O22, and O26 (O6, O16, and O26 correspond to O3, O13, and O23, respectively, in compound 2). 2 Average Ln-O bond distance of the fma oxygen atoms.
Table 5. Ln-Ln diagonal distances (in Å) and Ln-Ln-Ln angles (in °) in compounds 14.
Table 5. Ln-Ln diagonal distances (in Å) and Ln-Ln-Ln angles (in °) in compounds 14.
1 (La/Cl)2 (La/Br)3 (Eu/Cl)4 (Eu/Br)
Ln-Ln (Å)21.2721.3020.7420.85
18.0718.1517.7817.83
11.4311.5411.3611.47
Ln-Ln-Ln (°)86.1186.6187.3087.37
109.05109.15109.02109.48
159.92159.71159.18159.07
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Benmansour, S.; Hernández-Paredes, A.; Defez-Aznar, K.; Gómez-García, C.J. Two-Dimensional Lattices with Lanthanoids, Anilato Ligands and Formamide. Crystals 2023, 13, 939. https://doi.org/10.3390/cryst13060939

AMA Style

Benmansour S, Hernández-Paredes A, Defez-Aznar K, Gómez-García CJ. Two-Dimensional Lattices with Lanthanoids, Anilato Ligands and Formamide. Crystals. 2023; 13(6):939. https://doi.org/10.3390/cryst13060939

Chicago/Turabian Style

Benmansour, Samia, Antonio Hernández-Paredes, Kilian Defez-Aznar, and Carlos J. Gómez-García. 2023. "Two-Dimensional Lattices with Lanthanoids, Anilato Ligands and Formamide" Crystals 13, no. 6: 939. https://doi.org/10.3390/cryst13060939

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