Next Article in Journal
iT4SE-EP: Accurate Identification of Bacterial Type IV Secreted Effectors by Exploring Evolutionary Features from Two PSI-BLAST Profiles
Next Article in Special Issue
A Highly Selective Turn-On Fluorescent Probe for the Detection of Zinc
Previous Article in Journal
Applications of Hyaluronic Acid in Ophthalmology and Contact Lenses
Previous Article in Special Issue
3D Metal–Organic Frameworks Based on Co(II) and Bithiophendicarboxylate: Synthesis, Crystal Structures, Gas Adsorption, and Magnetic Properties
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Heteroleptic LaIII Anilate/Dicarboxylate Based Neutral 3D-Coordination Polymers

by
Olesya Y. Trofimova
1,
Arina V. Maleeva
1,
Irina V. Ershova
1,
Anton V. Cherkasov
1,
Georgy K. Fukin
1,
Rinat R. Aysin
2,
Konstantin A. Kovalenko
3 and
Alexandr V. Piskunov
1,*
1
G. A. Razuvaev Institute of Organometallic Chemistry of the Russian Academy of Sciences, Tropinin Str., 49, 603137 Nizhny Novgorod, Russia
2
A. N. Nesmeyanov Institute of Organometallic Chemistry of the Russian Academy of Sciences, Vavilova Str., 28, 119991 Moscow, Russia
3
Nikolaev Institute of Inorganic Chemistry, Siberian Branch of Russian Academy of Sciences, Acad. Lavrentiev Ave., 3, 630090 Novosibirsk, Russia
*
Author to whom correspondence should be addressed.
Molecules 2021, 26(9), 2486; https://doi.org/10.3390/molecules26092486
Submission received: 2 April 2021 / Revised: 22 April 2021 / Accepted: 22 April 2021 / Published: 24 April 2021

Abstract

:
Three new 3D metal–organic frameworks of lanthanum based on mixed anionic ligands, [(La2(pQ)2(BDC)4)·4DMF]n, [(La2(pQ)2(DHBDC)4)·4DMF]n, [(La2(CA)2(BDC)4)·4DMF]n (pQ—dianion of 2,5-dihydroxy-3,6-di-tert-butyl-para-quinone, CA—dianion of chloranilic acid, BDC-1,4-benzenedicarboxylate, DHBDC-2,5-dihydroxy-1,4-benzenedicarboxylate and DMF-N,N′-dimethylformamide), were synthesized using solvothermal methodology. Coordination polymers demonstrate the rare xah or 4,6T187 topology of a 3D framework. The homoleptic 2D-coordination polymer [(La2(pQ)3)·4DMF]n was obtained as a by-product in the course of synthetic procedure optimization. The thermal stability, spectral characteristics and porosity of coordination polymers were investigated.

Graphical Abstract

1. Introduction

Metal−organic frameworks (MOFs) and coordination polymers (CPs) are an emerging class of microporous solids, which have been intensively studied during the last two decades due to their structural and functional diversity. They have potential applications as gas adsorbents [1,2,3]; luminescent [4,5,6,7]; electrochemical or photophysical sensors [8,9]; and catalytic [10,11,12], optical [13], electrically conductive [14], and magnetic materials [15,16,17,18]. The physical and chemical properties, crystalline structure and topologies of coordination polymers depend on the properties of the organic ligands and/or the metal ions.
The most used types of ligands for MOFs’ construction are various di-, tri-, and tetracarboxylic acids [13]. Their use allows a variety of structures of different dimensions to be obtained, from linear to 3D frames. Among the unique modern trends is the combination of various anionic bridges to obtain heteroleptic polymeric systems. The combination of two or more types of ligands in one link leads to the manifestation of the unusual properties of MOFs and CPs [1,2,4,7]. For example, Fedin and coworkers [1,2] have shown that zinc complexes’ design based on dicarboxylic acids and diatomic alcohols allows the synthesis of functional coordination polymers with outstanding characteristics, excellent adsorption selectivity of benzene over cyclohexane, for example.
Another promising and rapidly developing ligand system in the construction of MOFs is a variety of anilates - derivatives of 2,5-dihydroxy-1,4-benzoquinone [18]. A unique feature of these ligands is their redox activity. Double-deprotonated anilates can exist in five redox states in complexes with metals. The introduction of different substituents in the 3 and 6 positions of the quinoid ring allows for a wide range of variations in their electronic and steric properties. Most commonly, researchers use chlorine [16,19,20], bromine [17,21,22], fluorine [23,24], nitro [25,26] and cyano-substituted [22,27,28] anilate ligands. The first paper on the use of 3,6-di-tert-butyl-2,5-dihydroxy-para-benzoquinone in the construction of 2D layered lanthanum MOFs was published [29]. Lanthanide MOFs/CPs are of great interest for their outstanding optical and magnetic properties [4,6,22,28,30,31,32,33]. They may appear as multifunctional materials for electro-optical, data storage, and sensing applications [5,27,34,35,36]. The first examples of heteroleptic NIR-Emitting YbIII/anilate-based coordination polymer nanosheets were prepared recently using terephthalic dicarboxylate coligands [7]. These 2D nanosheets are ideal objects for performing advanced photophysical studies by an innovative multiprobe approach. The research of perturbations of photoluminescence induced by different solvents, both aromatic and aliphatic, bearing electron-withdrawing and electro-donating groups, has demonstrated high-sensitivity to nitrobenzene [7].
In the present work, we report the synthesis and characterization of the first 3D-heteroleptic anilate/carboxylate/La(III) metal organic frameworks. These compound are formulated as [(La2(pQ)2(BDC)4)·4DMF]n (1), [(La2(pQ)2(DHBDC)4)·4DMF]n (2) and [(La2(CA)2(BDC)4)·4DMF]n (3) (pQ—dianionic form of 2,5-dihydroxy-3,6-di-tert-butyl-para-quinone, CA—dianionic form of chloranilic acid, BDC—1,4-benzenedicarboxylate, DHBDC—2,5-dihydroxy-1,4-benzenedicarboxylate, DMF—N,N′-dimethylformamide).

2. Results and Discussion

Compounds 13 were synthesized by a solvothermal technique in DMF (Scheme 1).
The synthetic procedure involves consecutive two-stage heating of the reaction mixture at 80 and 130 °C. This ensures a high yield of violet crystals of heteroleptic reaction products. The reaction under conditions of single-stage heating of the components in a DMF solution is accompanied by the formation of crystalline colorless lanthanum carboxylates and red homoleptic anilate derivatives as by-products. Additionally, an increase in the carboxylate ligand’s content in the initial reaction mixture contributes to an increase in the yield of the target product. Therefore, for polymers 1 and 2, the optimal ratio is H2pQ:H2bdc (H2dhbdc) = 1:4. For the chloranilic acid derivative 3, the ratio is H2CA:H2bdc = 1:3. It should be noted that the optimal ratio of reagents differs from the stoichiometric one. This is obviously due to the different rates of formation and crystallization of homoleptic polymers. Additionally, the anilate ligands have a stronger affinity toward the La(III) ion than the carboxylate one. This assertion was forecasted by the authors of [7]. The interaction Yb(III)/anilate/dicarboxylate with stoichiometric (2:1:1) ratio in the reaction produces coordination polymers with the 2:2.5:0.5 or 2:2:1 formula ratios [7]. A significant increase in the content of dicarboxylic acid in the reaction mixture makes it possible to obtain a polymer with a component ratio of La(III)/anilate/dicarboxylate—2:2:4, which dramatically affects the resulting CPs’ structures. In contrast to [7], it is possible to obtain three-dimensional structures instead of layered ones.
The La(III) CP (4) bearing pQ2− and DMF ligands only was prepared by the test synthesis excluding the dicarboxylate component (Scheme 2).
The SC XRD studies of 13 compounds were performed, revealing a series of new MOFs with similar structural compositions. MOFs 1 and 2 crystallize in the triclinic P-1 space group, while 3 crystallizes in the monoclinic P21/n. MOFs 1 and 2 display a rarely xah [37] topology of the 3D framework. The topology of the underlying network of 3 is 4,6T187 in the standard representation of the valence-bonded MOFs.
The simplest unit of 13 comprises two La(III) centers (Figure 1 and Figure 2). They are bridged by four carboxylate ligands forming the Chinese-lantern structure. Each La(III) center coordinates five oxygen atoms from four BDC2− (1, 3) or DHBDC2− (2) anions, two oxygen atoms from pQ2− (1, 2) or CA2− (3) anions, and two oxygen atoms of two DMF coordination molecules (Figure 3). Thus, the formal coordination number of La3+ in 13 is nine. The La…La distances in these dimeric units are 4.1978(4) Å in 1, 4.2239(2) Å in 2 and 4.1725(5) Å in 3.
Systematic analysis of the coordination geometries around the La(III) centers using SHAPE 2.1 software [38,39] reveals that the nine coordinated [40] lanthanide centers of complexes 1–3 adopt different geometries. The coordination polyhedron of lanthanum ion in compound 1 is best described as a spherical capped square antiprism (CSAPR-9, minimum CShM value is 1.383) (Figure 4a). The bottom and top edges are formed by O(2B), O(4A), O(5A), O(6) and O(1), O(3), O(5), O(7) atoms, respectively. The oxygen O(8) of one of the coordinated DMF molecules occupies the capped vertex. Coordination polyhedra in CPs 2 (Figure 4b) and 3 (Figure 4c) are quite close to the muffin (MFF-9, minimum CShM values are 0.926 and 0.868 for 2 and 3, respectively). The bottom base of these muffins include O(2B), O(4A), O(10) and O(2D), O(5C), O(7) atoms for 2 and 3, respectively. Apical vertexes depicted by oxygens O(3) (compound 2) and O(6B) (compound 3). Pentagonal planes consist of O(1), O(6A), O(7A), O(6), O(9) and O(1), O(3), O(4A), O(6E), O(8) atoms for 2 and 3, respectively.
A large excess of dicarboxylic acid contributes to the formation of layers constructed from carboxylate lanterns (Figure 5, Figure 6 and Figure 7). The latter form two-dimensional grids with four-membered cells. The structures of polymers 1 and 2 are close to each other. The carboxylate layers in them are almost flat. In derivative 3, a zigzag formation of the layers is observed, which leads to its thickening. The layers in 13 are cross-linked into a three-dimensional structure through anilate ligands. This mechanism of a three-dimensional framework formation explains the difference between the here presented results and those obtained in the preparation of heteroleptic ytterbium complexes [7]. In the case of an excess of the anilic acid, a typical for such complexes (6,3)-2D topology of six-membered rings with hexagonal cavities is formed on the first stage, and dicarboxylate ligands replace some anilate ligands in this lattice. Thus, the predominance of anilic over dicarboxylate ligands in the reaction mixture forms a two-dimensional coordination polymer structure. At the same time, the inverse ratio of the reagents leads to a 3D framework.
The dianionic form of anilate ligands can bind to metals either as a bridging (bis)bidentate ligand, which has a dianion-like structure (Scheme 3, type I and II), or as a terminal bidentate ligand with an ortho-quinoid structure (type III) [41,42]. The dianions of 2,5-dihydroxy-3,6-di-tert-butyl-para-quinone in 12 and chloranilic acid in 3 consist of two delocalized π-electron systems connected by single C-C bonds (1.552(3) Å). Other C-C distances of six-membered cycle and C-O bonds lies in the narrow ranges 1.404(3)–1.409(3) Å and 1.267(2)–1.271(2) Å, respectively. Such delocalization of electron density in the quinoid ring is characteristic of the bridging mode of anilate derivatives.
Ligands BDC2− in 1 and 3 or DHBDC2– in 2 adopted bridging mode, linking four La3+ ions. The types of bridging are shown in Scheme 4. In compounds 1 and 2, two out of four dicarboxylate ligands act as a linker with each C(O)O-group coordinated by μ21: κ1-type (type a in Scheme 4), while in the other two ligands, the functional groups have the μ21: κ2 coordination mode (type b in Scheme 4) (Figure 3a,b). For all four dicarboxylate ligands in 3, the μ21: κ2 coordination mode for one C(O)O-group and coordination by μ21: κ1-type for another one were observed (type c in Scheme 4, Figure 3c). The La-Odicarb bond distances in 13 are in the range of 2.466(3)–2.849(2) Å, which are similar to those of other related La(III) compounds with dicarboxylate ligands [43,44]. Selected bond lengths are presented in Table 1.
Symmetry transformations used to generate equivalent atoms:
1: 
(A) −x + 1, −y + 1, −z + 1; (B) −x + 1, −y + 1, −z; (C) −x, −y + 2, −z + 1; (D) −x + 2, −y + 1, −z + 1
2: 
(A) −x + 1, −y + 1, −z + 1; (B) −x + 1, −y + 1, −z; (C) −x + 1, −y, −z + 1; (D) −x, −y + 1, −z + 1
3: 
(A) −x + 3, −y, −z + 2; (B) −x + 5/2, y + 1/2, −z + 3/2; (C) x + 1/2, −y−1/2, z + 1/2; (D) −x + 2, −y, −z + 2; (E) x + 1/2, −y −1/2, z + 1/2
4: 
(A) −x + 1, −y + 2, −z + 2; (B) −x + 1, −y + 1, −z + 2; (C) −x, −y + 1, −z + 1.
According to X-ray structural analysis, compounds 2 and 3 contain free accessible voids filled with solvent molecules. The volume of these voids for compound 2 was estimated at 13.8% of unit cell volume (Figure 7a). The voids volume in 3 is about 15.6% (Figure 7b) [45]. Visually, compounds 2 and 3 lost crystallinity within a day, which is associated with the removal of guest DMF molecules from the pores of compounds. The data of elemental and thermogravimetric analyses confirm the absence of guest DMF molecules in the dried samples of 2 and 3.
Metal–organic coordination polymer 4 crystallizes in the triclinic P-1 space group. The topology of the underlying net of 4 is hcb and typical for this type of compound [7,21,22,33]. The simplest unit of [(La2(pQ)3)∙4DMF]n comprises two La(III) cations, three pQ2− dianions and four coordinated N,N′-dimethylformamide ligands (Figure 8). The atom La(III) is 8-coordinated with a distorted triangular dodecahedron geometry (TDD-8, minimum CShM value is 1.148) (Figure 8). The metal–organic coordination polymer 4 has a similar structure to the recently published derivatives [Ln2(pQ)3∙4DMAA]n (Ln = La, Pr, Nd, DMAA = N,N′-dimethylacetamide) [29]. The main difference is the type of anilate ligands binding to lanthanide ions (Scheme 3). For complexes with dimethylacetamide, two of the three ligands are characterized by type I coordination, and the third one has p-quinoid type II coordination mode. In contrast, all three dianions of 2,5-dihydroxy-3,6-di-tert-butyl-para-quinone in 4 are coordinated by type I and characterized by two delocalized π-electron systems (C-O 1.28(2)-1.32(2) Å) connected by single C–C bonds (1.54(2)-1.551(5) Å). Selected bond lengths in coordination polymer 4 are presented in Table 1.
Thermal stability of obtained compounds 13 was examined by TGA and DTA (Figure 9). Compounds 13 started to decompose above 150 °C and showed a slight mass decrease upon heating to 260 °C due to a continuous loss of DMF molecules. The observed mass loss was 10%, 10%, and 8% for 1, 2, and 3, respectively, which corresponds to 1.5 DMF molecules. Up to 150 °C, no weight loss was observed on the TGA curve, which indicates the absence of occluded solvent and guest DMF molecules in the samples. According to DTA data, heating above 270 °C leads to several different processes, which lead to the decomposition of the framework of compounds 13. Compound 4 demonstrates thermal stability up to 170 °C. The sample starts to decompose above 170 °C and shows a steady mass decrease upon heating to 250 °C due to a continuous loss of coordination solvent molecules. The observed mass loss equal to 11% corresponds to two DMF molecules in counting on the simplest unit of [(La2(pQ)3)∙4DMF]. Decomposition of the framework occurred at 298 °C. TG curves for 4 are presented in Figure S2 (ESI).
The porous structure of compounds 13 was investigated using the example of compound 1, because this compound retains its crystallinity over a prolonged time period. An analysis of the porous structure was performed by a carbon dioxide adsorption technique at 195 K. Initially, the compound 1 was activated in a dynamic vacuum using the standard “outgas” option of the equipment at 453 K for 2 or 6 h. Measured adsorption isotherms of 1 are represented in Figure 10. Compound 1 has low porosity, pore volume and specific surface area. Longer activation leads to a decrease in the adsorption capacity, as evidenced by a drop in the adsorbed volume of CO2. Evidently, this is due to the destruction of the 3D structure when the molecules of the coordinated DMF are eliminated. This is also indicated by a significant deterioration of the powder diffractogram after the completion of the sorption–desorption cycle. Calculated parameters of the porous structure are shown in Table 2. Pore size distributions were calculated using the DFT approach, which shows good agreement between the measured and calculated isotherms and is presented in ESI. Isotherms are typical for microporous materials with slit pores, which was also proven by DFT PSD calculations.

3. Materials and Methods

3.1. Reagents and Methods

All reactants were purchased from Sigma Aldrich. 2,5-dihydroxy-3,6-di-tert-butyl-p-benzoquinone was synthesized according to the previously reported procedure [46]. All synthetic manipulations were performed under Schlenk line conditions. Solvents were purified by standard methods [47]. Elemental analyses were performed with an Elementar Vario El cube instrument. Electronic absorption (UV-vis) spectra in the range 200–900 nm of nujol mulls were recorded on a Carl Zeiss Jena Specord M400 spectrophotometer. IR-spectra of the studied compounds were recorded on a FSM1201 Fourier-IR spectrometer in a nujol using KBr plates in the range 4000–400 cm−1. TGA of compound 13 was measured on a Shimadzu DTG-60H synchronous thermal analyzer instrument from 30 to 600 °C using an Al pan and heated at a rate of 5 °C/min under an air atmosphere. TGA of compound 4 was measured on a Mettler Toledo TGA/DSC3+ instrument from 30 to 500 °C using an PCA pan and heated at a rate of 5 °C/min under a nitrogen atmosphere.
The X-ray data for 14 were collected with Bruker D8 Quest (13) and Rigaku OD Xcalibur (4) diffractometers (Mo-radiation, ω-scans technique, λ = 0.71073Å, T = 100 K) using APEX3 [48] and CrysAlisPro [49] software packages. The structures were solved via an intrinsic phasing algorithm and refined by full-matrix least squares against F2 using SHELX [50,51]. SADABS [52] and scaling algorithms implemented in CrysAlisPro were used to perform absorption corrections. All non-hydrogen atoms in 14 were found from Fourier syntheses of electron density and refined anisotropically. All hydrogen atoms were placed in calculated positions and refined isotropically in the “riding” model with U(H)iso = 1.2Ueq of their parent atoms (U(H)iso = 1.5Ueq for methyl groups). The crystal data and structure refinement details for compounds 14 are presented in Table 3.
One solvate DMF molecule in the asymmetric unit of 2 was modelled by SQUEEZE/Platon [53]. Coordination polymer 4 was refined as 2-component non-merohedral twin with a domain ratio 0.65/0.35.
The crystallographic data and structure refinement details are shown in Table S1. CCDC 2,074,328 (1), 2,074,329 (2), 2,074,330 (3) and 2,074,331 (4) contain the supplementary crystallographic data for this paper. These data are provided free of charge by The Cambridge Crystallographic Data Centre: ccdc.cam.ac.uk/structures.
An analysis of the porous structure was performed by a carbon dioxide adsorption technique using Quantochrome’s Autosorb iQ at 195 K. Cryostat CryoCooler was used to adjust temperature with 0.05 K accuracy. Carbon dioxide adsorption−desorption isotherms were measured within the range of relative pressures from 10 to 3 till 0.995. The specific surface area was calculated from the data obtained on the basis of the conventional BET, Langmuir and DFT models. Pore size distributions were calculated using the DFT method. The database of the National Institute of Standards and Technology available at http://webbook.nist.gov/chemistry/fluid/ (accessed on 23 April 2021) was used as a source of p−V−T relations at experimental pressures and temperatures.
X-ray powder diffraction measurements were performed on a Shimadzu LabX XRD-6100 X-ray Powder Diffractometer. Storage of crystalline compounds after their separation from the mother liquors led to a partial loss of solvate DMF molecules, which led to the destruction of the crystals studied by X-ray diffraction and the formation of a new crystalline phase. Apparently, the formation of this new phase is reflected in the appearance of extra peaks in the powder diffractogram.

3.2. Synthesis

3.2.1. Synthesis of 1

[(La2(pQ)2(BDC)4)∙4DMF]n (1). A mixture of LaCl3∙7H2O (0.04 mmol), H2BDC (0.16 mmol) and H2pQ (0.04 mmol) in a mole ratio of 1:4:1 in DMF (2 mL) was heated sequentially at 80 °C for 24 h and 130 °C for 48 h in a sealed glass ampule. The obtained purple crystals were separated by the filtration, washed with DMF (2 × 2 mL) and dried on air. Yield was 90% based on LaCl3∙7H2O. Elemental analysis calculated for C42H54N4O16La2: C, 43.92; H, 4.74; N, 4.88. Found: C, 43.45; H, 4.62; N 4.81%. IR (Nujol, KBr, cm–1): 1671 w, 1648 w, 1607 w, 1587 w, 1489 w, 1347 w, 1308 w, 1257 m, 1219 m, 1200 m, 1154 m, 1107 w, 1086 s, 1065 m, 1047 s, 1018 m, 972 s, 928 s, 903 m, 887 m, 866 s, 841 w, 810 m, 796 m, 760 w, 750 w, 673 w, 656 m, 511 w, 484 w. Electronic absorption spectrum (Nujol, λ (nm)): 336, 382, 508.

3.2.2. Synthesis of 2

[(La2(pQ)2(DHBDC)4)∙4DMF]n (2). A mixture of LaCl3∙7H2O (0.04 mmol), H2DHBDC (0.12 mmol) and H2pQ (0.04 mmol) in a mole ratio of 1: 3: 1 in DMF (2 mL) was heated sequentially at 80 °C for 24 h and 130 °C for 48 h in a sealed glass ampule. The obtained purple crystals were separated by the filtration, washed with DMF (2 × 2 mL), and dried on air. Yield was 81% based on LaCl3∙7H2O. Visually, the compound lost its crystallinity within 1 day. According to the data of X-ray structural analysis, compound 2 contains 2 DMF molecules per simplest unit. According to elemental analysis, after the loss of crystallinity, the compound does not contain guest solvent. Elemental analysis calculated for [C24H34LaN3O11]: C, 41.60; H, 4.49; N, 4.62. Found: C, 41.20; H, 4.32; N 4.57%. IR (Nujol, KBr, cm–1): 3130 w, 1656 w, 1644 w, 1605 w, 1584 w, 1484 w, 1364 m, 1342 w, 1235 w, 1105 w, 1058 s, 1047 s, 1019 s, 970 s, 925 s, 896 s, 873 m, 862 w, 819 w, 813 w, 794 w, 784 w, 733 w, 673 w, 660 m, 617 s, 607 m, 549 w, 542 w, 509 w, 482 w. Electronic absorption spectrum (Nujol, λ (nm)): 350, 390, 520–540.

3.2.3. Synthesis of 3

[(La2(CA)2(BDC)4)∙4DMF]n (3). A mixture of LaCl3∙7H2O (0.04 mmol), H2BDC (0.12 mmol) and H2CA (0.04 mmol) in a mole ratio of 1:3:1 in DMF (2 mL) was heated sequentially at 80 °C for 24 h and 130 °C for 48 h in a sealed glass ampule. The obtained purple crystals were separated by the filtration, washed with DMF (2 × 2 mL) and dried on air. Yield was 81% based on LaCl3∙7H2O. Visually, the compound lost its crystallinity within 1 day. According to the data of X-ray structural analysis, compound 3 contains 0.6 DMF molecules per simplest unit. According to elemental analysis, after the loss of crystallinity, the compound does not contain guest solvent. Elemental analysis calculated for [C34H36Cl4N4O16La2]: C, 36.94; H, 3.28; N, 5.07. Found: C, 37.16; H, 3.22; N 5.39%. IR (Nujol, KBr, cm–1): 1676 w, 1654 w, 1604 w, 1587 w, 1512 w, 1439 w, 1418 w, 1312 m, 1296 m, 1252 s, 1151 m, 1130 s, 1106 w, 1164 s, 1015 m, 993 s, 890 s, 862 s, 841 w, 826 m, 816 m, 785 s, 754 w, 675 w, 595 m, 576 w, 540 s, 510 w. Electronic absorption spectrum (Nujol, λ (nm)): 350, 520.

3.2.4. Synthesis of 4

[(La2(pQ)3∙4DMF]n. A mixture of solid LaCl3·7H2O (0.1 mmol) and H2pQ (0.1 mmol) was placed in a glass ampoule, and N,N′-dimethylformamide (5 mL) was added. Ampoule was evacuated, sealed and heated at 130 °C for 1 day. The obtained burgundy crystals were separated by the filtration, washed twice by 5 mL of N,N′-dimethylformamide and dried on air. Yield was 90% based on H2pQ. C27H41LaN2O8. Calculated C, 49.10; H, 6.26; N 4.24%. Found C, 49.51; H, 6.20; N 4.34%. IR (Nujol, KBr) cm−1: 1661 w, 1592 m, 1536 w, 1472 w, 1445 w, 1379 w, 1339 w, 1254 m, 1211 m, 1202 m, 1105 w, 1059 m, 1047 w, 1013 s, 964 m, 928 s, 900 w, 865 s, 793 m, 725 s, 673 w, 659 w, 621 m, 480 w. Electronic absorption spectrum (Nujol, λ (nm)): 352, 540.

4. Conclusions

In summary, we reported the first examples of lanthanide heteroleptic anilate/dicarboxylate three-dimensional metal–organic frameworks constructed from mixed dianionic ligands. The synthesis was carried out under conditions of an excess of dicarboxylic acid. This led to the assembly of two-dimensional layers formed by binuclear lanthanum carboxylate lanterns, spliced into three-dimensional structures by anilate crosslinks. The synthesized structures have low porosity and are thermally stable up to a temperature of 150 °C. We believe that the synthetic approach developed in this work will be particularly useful and can be used for the purposeful construction of functional MOFs based on the entire range of lanthanides.

Supplementary Materials

The following are available online, Figure S1: Pore size distribution curves for compounds 1; Figure S2: Pore size distribution curves for TG curve for 4; Figure S3: PXRD patterns for 14; Figure S4: The molecular structure of the binuclear unit in 2. Shape analysis for 13, Shape analysis for 4.

Author Contributions

Conceptualization, supervision, writing—review and editing, A.V.P.; writing—original draft preparation O.Y.T.; synthesis and methodology O.Y.T., A.V.M. and I.V.E.; formal analysis and discussion A.V.C., G.K.F., R.R.A. and K.A.K. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by Russian Foundation for Basic Research No. 18-29-04041_mk.

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 X-ray diffraction experiments and TGA/DTA were performed using the equipment from the “Analytical Center of the IOMC RAS”. The PXRD research was carried out with the equipment of the Collective Usage Center “New Materials and Resource-saving Technologies” (Lobachevsky State University of Nizhny Novgorod).

Conflicts of Interest

The authors declare no conflict of interest.

Sample Availability

Samples of compounds 14 are available from the authors.

References

  1. 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]
  2. 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]
  3. Adil, K.; Belmabkhout, Y.; Pillai, R.S.; Cadiau, A.; Bhatt, P.M.; Assen, A.H.; Maurinb, G.; Eddaoudi, M. Gas/vapour separation using ultra-microporous metal–organic frameworks: Insights into the structure/separation relationship. Chem. Soc. Rev. 2017, 46, 3402–3430. [Google Scholar] [CrossRef]
  4. Yao, X.; Wang, X.; Han, Y.; Yan, P.; Li, Y.; Li, G. Structure, color-tunable luminescence, and UV-vis/NIR benzaldehyde detection of lanthanide coordination polymers based on two fluorinated ligands. CrystEngComm 2018, 20, 3335–3343. [Google Scholar] [CrossRef]
  5. Ren, Y.-X.; Zhao, X.-L.; Wang, Z.-X.; Pan, Y.; Li, H.-P.; Wang, F.-Y.; Zhu, S.-F.; Shao, C.-H. Effects of template molecules on the structures and luminescence intensities of a series of porous Tb-MOFs based on the 2-nitroterephthalate ligand. RSC Adv. 2018, 8, 17497–17503. [Google Scholar] [CrossRef] [Green Version]
  6. Zhang, X.; Zhan, Z.; Liang, X.; Chen, C.; Liu, X.; Jia, Y.; Hu, M. Lanthanide-MOFs constructed from mixed dicarboxylate ligands as selective multi-responsive luminescent sensors. Dalton Trans. 2018, 47, 3272–3282. [Google Scholar] [CrossRef] [PubMed]
  7. 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]
  8. D’Alessandro, D.M. Exploiting redox activity in metal–organic frameworks: Concepts, trends and perspectives. Chem. Commun. 2016, 52, 8957–8971. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  9. Liu, L.; DeGayner, J.A.; Sun, L.; Zee, D.Z.; Harris, T.D. Reversible redox switching of magnetic order and electrical conductivity in a 2D manganese benzoquinoid framework. Chem. Sci. 2019, 10, 4652–4661. [Google Scholar] [CrossRef] [Green Version]
  10. Burnett, D.L.; Oozeerally, R.; Pertiwi, R.; Chamberlain, T.W.; Cherkasov, N.; Clarkson, G.J.; Krisnandi, Y.K.; Degirmenci, V.; Walton, R.I. A hydrothermally stable ytterbium metal–organic framework as a bifunctional solid-acid catalyst for glucose conversion. Chem. Commun. 2019, 55, 11446–11449. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  11. Hu, Z.; Zhao, D. Metal–organic frameworks with Lewis acidity: Synthesis, characterization, and catalytic applications. CrystEngComm 2017, 19, 4066–4081. [Google Scholar] [CrossRef]
  12. Wang, L.; Li, W.-X.; Gu, X.-M.; Zhang, W.-L.; Ni, L. Structure Variation from One-Dimensional Chain to Three-Dimensional Architecture: Effect of Ligand on Construction of Lanthanide Coordination Polymers. J. Chem. Sci. 2017, 129, 271–280. [Google Scholar] [CrossRef] [Green Version]
  13. Li, B.; Wen, H.-M.; Cui, Y.; Zhou, W.; Qian, G.; Chen, B. Emerging Multifunctional Metal–Organic Framework Materials. Adv. Mater. 2016, 28, 8819–8860. [Google Scholar] [CrossRef]
  14. Chen, G.; Gee, L.B.; Xu, W.; Zhu, Y.; Lezama-Pacheco, J.S.; Huang, Z.; Li, Z.; Babicz, J.T.; Choudhury, J.S.; Chang, T.-H.; et al. Valence-Dependent Electrical Conductivity in a 3D Tetrahydroxyquinone-Based Metal−Organic Framework. J. Am. Chem. Soc. 2020, 142, 21243–21248. [Google Scholar] [CrossRef] [PubMed]
  15. 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]
  16. Chen, J.; Sekine, Y.; Komatsumaru, Y.; Hayami, S.; Miyasaka, H. Thermally Induced Valence Tautomeric Transition in a Two-Dimensional Fe-Tetraoxolene Honeycomb Network. Angew. Chem. Int. Ed. 2018, 57, 12043–12047. [Google Scholar] [CrossRef]
  17. Martínez-Hernández, C.; Gómez-Claramunt, P.; Benmansour, S.; Gómez-García, C.J. Pre- and post-syntheticmodulation of the ordering temperatures in a family of anilato-based magnets. Dalton Trans. 2019, 48, 13212–13223. [Google Scholar] [CrossRef] [PubMed]
  18. Kitagawa, S.; Kawata, S. Coordination compounds of 1,4-dihydroxybenzoquinone and its homologues. Structures and properties. Coord. Chemi. Rev. 2002, 224, 11–34. [Google Scholar] [CrossRef]
  19. Bondaruk, K.; Hua, C. Effect of Counterions on the Formation and Structures of Ce(III) and Er(III) Chloranilate Frameworks. Cryst. Growth Des. 2019, 19, 3338–3347. [Google Scholar] [CrossRef]
  20. AndrosˇDubraja, L.; Molcanov, K.; Zilic, D.; Kojic-Prodic, B.; Wenger, E. Multifunctionality and size of the chloranilate ligand define the topology of transition metal coordination polymers. New J. Chem. 2017, 41, 6785–6794. [Google Scholar] [CrossRef]
  21. Martínez-Hernández, C.; Benmansour, S.; García, C.J.G. Modulation of the ordering temperature in anilato-based magnets. Polyhedron 2019, 170, 122–131. [Google Scholar] [CrossRef]
  22. Hernández-Paredes, A.; Cerezo-Navarrete, C.; García, C.J.G.; Benmansour, S. Slow relaxation in doped coordination polymers and dimers based on lanthanoids and anilato ligands. Polyhedron 2019, 170, 476–485. [Google Scholar] [CrossRef]
  23. 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]
  24. 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]
  25. Simonson, A.N.; Kareis, C.M.; Ovanesyan, N.S.; Baumann, D.O.; Rheingold, A.L.; Arif, A.M.; Miller, J.S. Seven-coordinate tetraoxolate complexes. Polyhedron 2018, 139, 215–221. [Google Scholar] [CrossRef]
  26. 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]
  27. 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]
  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. 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]
  30. Abrahams, B.F.; Coleiro, J.; Ha, K.; Hoskins, B.F.; Orchard, S.D.; Robson, R. Dihydroxybenzoquinone and chloranilic acid derivatives of rare earth metals. Dalton Trans. 2002, 8, 1586–1594. [Google Scholar] [CrossRef]
  31. Luo, J.; Liu, B.S.; Cao, C.; Wei, F. Neodymium(III) organic frameworks (Nd-MOF) as near infrared fluorescent probe for highly selectively sensing of Cu2+. Inorg. Chem. Comm. 2017, 76, 18–21. [Google Scholar] [CrossRef]
  32. Li, C.Y.; Cai, D.M.; Yin, J.C.; Cai, L.P.; Zeng, M.; Wang, J.; Zhu, W.H. Crystal Structure, Fluorescence Spectroscopy, and Electrochemical Property of Two Neodymium Coordination Polymers with Phenoxy Acids. Russ. J. Coord. Chem. 2016, 42, 476–485. [Google Scholar] [CrossRef]
  33. Benmansour, S.; Gómez-García, C.J. Lanthanoid-Anilato Complexes and Lattices. Magnetochemistry 2020, 6, 71. [Google Scholar] [CrossRef]
  34. Wang, Y.; Liu, X.; Li, X.; Zhai, F.; Yan, S.; Liu, N.; Chai, Z.; Xu, Y.; Ouyang, X.; Wang, S. Direct Radiation Detection by a Semiconductive Metal−Organic Framework. J. Am. Chem. Soc. 2019, 141, 8030–8034. [Google Scholar] [CrossRef]
  35. Artizzu, F.; Atzori, M.; Liu, J.; Mara, D.; Hecke, K.V.; Deun, R.V. Solution-processable Yb/Er 2D-layered metallorganic frameworks with high NIR-emission quantum yields. J. Mater. Chem. C 2019, 7, 11207–11214. [Google Scholar] [CrossRef]
  36. 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]
  37. Alexandrov, E.V.; Blatov, V.A.; Kochetkov, A.V.; Proserpio, D.M. Underlying nets in three-periodic coordination polymers: Topology, taxonomy and prediction from a computer-aided analysis of the Cambridge Structural Database. CrystEngComm 2011, 13, 3947–3958. [Google Scholar] [CrossRef]
  38. Alvarez, S.; Alemany, P.; Casanova, D.; Cirera, J.; Llunell, M.; Avnir, D. Shape maps and polyhedral interconversion paths in transition metal chemistry. Coord. Chem. Rev. 2005, 249, 1693–1708. [Google Scholar] [CrossRef]
  39. Llunell, M.; Casanova, D.; Cirera, J.; Alemany, P.; Alvarez, S. SHAPE (2.1); Universitat de Barcelona: Barcelona, Spain, 2013. [Google Scholar]
  40. Ruiz-Martinez, A.; Casanova, D.; Alvarez, S. Polyhedral Structures with an Odd Number ofVertices: Nine-Coordinate Metal Compounds. Chem. Eur. J. 2008, 14, 1291–1303. [Google Scholar] [CrossRef] [PubMed]
  41. 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]
  42. Vuković, V.; Molcanov, K.I.; Jelsch, C.; Wenger, E.; Krawczuk, A.; Jurić, M.; Dubraja, L.A.; Kojić-Prodić, B. Malleable Electronic Structure of Chloranilic Acid and Its Species Determined by X-ray Charge Density Studies. Cryst. Growth Des. 2019, 19, 2802–2810. [Google Scholar] [CrossRef]
  43. 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]
  44. Wang, Y.-L.; Jiang, Y.-L.; Xiahou, Z.-J.; Fu, J.-H.; Liu, Q.-Y. Diversity of lanthanide(III)-2,5-dihydroxy-1,4-benzenedicarboxylate extended frameworks: Syntheses, structures, and magnetic properties. Dalton Trans. 2012, 41, 11428–11437. [Google Scholar] [CrossRef] [PubMed]
  45. Barbour, L.J. Crystal porosity and the burden of proof. Chem. Commun. 2006, 1163–1168. [Google Scholar] [CrossRef] [PubMed]
  46. 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]
  47. Perrin, D.D.; Armarego, W.L.F.; Perrin, D.R. Purification of Laboratory Chemicals; Pergamon: Oxford, UK, 1980. [Google Scholar]
  48. Bruker. APEX3; Bruker AXS Inc.: Madison, WI, USA, 2018. [Google Scholar]
  49. Rigaku Oxford Diffraction; CrysAlisPro Ver. 1.171.37.35; Rigaku Corporation: Wroclaw, Poland, 2014.
  50. Sheldrick, G.M. SHELXT—Integrated space-group and crystalstructure determination. Acta Cryst. 2015, A71, 3–8. [Google Scholar]
  51. Sheldrick, G.M. Crystal structure refinement with SHELXL. Acta Cryst. C 2015, 71, 3–8. [Google Scholar] [CrossRef] [PubMed]
  52. Krause, L.; Herbst-Irmer, R.; Sheldrick, G.M.; Stalke, D. Comparison of silver and molybdenum microfocus X-ray sources for single-crystal structure determination. J. Appl. Cryst. 2015, 48, 3–10. [Google Scholar] [CrossRef] [Green Version]
  53. Spek, A.L. PLATON SQUEEZE: A tool for the calculation of the disordered solvent contribution to the calculated structure factors. Acta Cryst. C 2015, 71, 9–18. [Google Scholar] [CrossRef] [Green Version]
Scheme 1. Synthetic scheme for the preparation of coordination polymers 13.
Scheme 1. Synthetic scheme for the preparation of coordination polymers 13.
Molecules 26 02486 sch001
Scheme 2. Synthetic scheme for the preparation of coordination polymer 4.
Scheme 2. Synthetic scheme for the preparation of coordination polymer 4.
Molecules 26 02486 sch002
Figure 1. The molecular structure of the binuclear unit in 1. Color code: La, light blue; C, gray; Cl, green; O, red; N, violet.
Figure 1. The molecular structure of the binuclear unit in 1. Color code: La, light blue; C, gray; Cl, green; O, red; N, violet.
Molecules 26 02486 g001
Figure 2. The molecular structure of the crystallographic unit in 3. Color code: La, light blue; C, gray; Cl, green; O, red; N, violet.
Figure 2. The molecular structure of the crystallographic unit in 3. Color code: La, light blue; C, gray; Cl, green; O, red; N, violet.
Molecules 26 02486 g002
Figure 3. The core of the binuclear units of 1 (a), 2 (b) and 3 (c). Color code: La, light blue; C, gray; O, red.
Figure 3. The core of the binuclear units of 1 (a), 2 (b) and 3 (c). Color code: La, light blue; C, gray; O, red.
Molecules 26 02486 g003
Figure 4. The coordination polyhedron of La cations in 1 (a), 2 (b) and 3 (c).
Figure 4. The coordination polyhedron of La cations in 1 (a), 2 (b) and 3 (c).
Molecules 26 02486 g004
Figure 5. The fragment of crystal packing of 1 along the (100) (left) and (001) (right) vectors. DMF molecules are omitted for clarity. Color code: La, light blue; anilic ligands, red; dicarboxylate ligands, blue.
Figure 5. The fragment of crystal packing of 1 along the (100) (left) and (001) (right) vectors. DMF molecules are omitted for clarity. Color code: La, light blue; anilic ligands, red; dicarboxylate ligands, blue.
Molecules 26 02486 g005
Figure 6. The fragment of crystal packing of 3 along the (100) (left) and (101) (right) vectors. DMF molecules are omitted for clarity. Color code: La, light blue; anilic ligands, red; dicarboxylate ligands, blue.
Figure 6. The fragment of crystal packing of 3 along the (100) (left) and (101) (right) vectors. DMF molecules are omitted for clarity. Color code: La, light blue; anilic ligands, red; dicarboxylate ligands, blue.
Molecules 26 02486 g006
Figure 7. The voids in crystal packing of 2 along the (001) vector (a) and 3 along the (100) vector (b). The voids volumes were computed with probe radius 1.2 Å and approx. grid spacing 0.7 Å. Color code: the outer side of the voids, red; the inner side, yellow.
Figure 7. The voids in crystal packing of 2 along the (001) vector (a) and 3 along the (100) vector (b). The voids volumes were computed with probe radius 1.2 Å and approx. grid spacing 0.7 Å. Color code: the outer side of the voids, red; the inner side, yellow.
Molecules 26 02486 g007
Scheme 3. Three different coordination modes of the anilate-type ligand in metal complexes.
Scheme 3. Three different coordination modes of the anilate-type ligand in metal complexes.
Molecules 26 02486 sch003
Scheme 4. The coordination modes of dicarboxylate ligands (a,b) in compounds 1 and 2, and (c) in 3.
Scheme 4. The coordination modes of dicarboxylate ligands (a,b) in compounds 1 and 2, and (c) in 3.
Molecules 26 02486 sch004
Figure 8. Molecular structure of binuclear unit of 4 (left); crystal structure of 4 along the (100) vector (right). Color code: La, light blue; C, gray; Cl, green; O, red; N, violet.
Figure 8. Molecular structure of binuclear unit of 4 (left); crystal structure of 4 along the (100) vector (right). Color code: La, light blue; C, gray; Cl, green; O, red; N, violet.
Molecules 26 02486 g008
Figure 9. TG curves for 1 (blue line), 2 (green line) and 3 (red line). DTA curves are shown with dash lines.
Figure 9. TG curves for 1 (blue line), 2 (green line) and 3 (red line). DTA curves are shown with dash lines.
Molecules 26 02486 g009
Figure 10. CO2 adsorption (, ▪)-desorption () isotherms at 195 K on 1 activated during 2 and 6 h s.
Figure 10. CO2 adsorption (, ▪)-desorption () isotherms at 195 K on 1 activated during 2 and 6 h s.
Molecules 26 02486 g010
Table 1. Selected bond lengths (Å) in 14.
Table 1. Selected bond lengths (Å) in 14.
Bond1Bond2∙2DMF
La(1)-O(1)2.477(2)La(1)-O(1)2.460(2)
La(1)-O(2B)2.516(2)La(1)-O(2B)2.505(2)
La(1)-O(3)2.513(2)La(1)-O(3)2.495(2)
La(1)-O(4A)2.510(2)La(1)-O(4A)2.48(2)
La(1)-O(5)2.582(2)La(1)-O(6)2.514(2)
La(1)-O(5A)2.731(2)La(1)-O(6A)2.849(2)
La(1)-O(6A)2.608(2)La(1)-O(7A)2.518(5)
La(1)-O(7)2.568(2)La(1)-O(9)2.569(2)
La(1)-O(8)2.580(5)La(1)-O(10)2.496(6)
O(1)-C(1)1.271(2)O(1)-C(1)1.267(2)
O(2)-C(3)1.267(2)O(2)-C(3)1.259(2)
C(1)-C(2)1.409(3)C(1)-C(2)1.402(3)
C(1)-C(3B)1.552(3)C(1)-C(3B)1.540(3)
C(2)-C(3)1.404(3)C(2)-C(3)1.411(3)
O(3)-C(8)1.262(2)O(3)-C(8)1.263(4)
O(4)-C(8)1.263(2)O(4)-C(8)1.319(5)
C(8)-C(9)1.511(2)C(8)-C(9)1.499(4)
C(9)-C(10)1.388(3)C(9)-C(10)1.390(4)
C(9)-C(11C)1.390(3)C(9)-C(11C)1.35(2)
C(10)-C(11)1.389(3)C(10)-C(11)1.393(5)
O(5)-C(12)1.245(2)O(6)-C(12)1.275(3)
O(6)-C(12)1.255(2)O(7)-C(12)1.261(4)
C(12)-C(13)1.501(3)C(12)-C(13)1.486(4)
C(13)-C(15D)1.396(3)C(13)-C(14)1.407(4)
C(13)-C(14)1.398(3)C(15)-C(13)1.391(4)
C(14)-C(15)1.390(3)C(15)-C(14D)1.654(6)
Bond3∙0.6DMFBond4
La(1)-O(1)2.58(2)La(1)-O(1)2.463(7)
La(1)-O(2D)2.572(7)La(1)-O(2A)2.458(7)
La(1)-O(3)2.474(3)La(1)-O(3)2.488(6)
La(1)-O(4A)2.466(3)La(1)-O(4C)2.503(6)
La(1)-O(5C)2.567(3)La(1)-O(5)2.488(6)
La(1)-O(6B)2.501(3)La(1)-O(6B)2.495(6)
La(1)-O(6E)2.729(3)La(1)-O(7)2.52(2)
La(1)-O(7)2.511(6)La(1)-O(8)2.53(2)
La(1)-O(8)2.494(7)O(1)-C(1)1.29(2)
O(1)-C(1)1.253(6)O(2)-C(3)1.28(2)
O(2)-C(3)1.251(6)C(1)-C(2)1.378(7)
C(1)-C(2)1.398(7)C(1)-C(3A)1.54(2)
C(1)-C(3D)1.58(2)C(2)-C(3)1.377(7)
C(2)-C(3)1.414(7)O(3)-C(8)1.288(9)
O(3)-C(4)1.256(6)O(4)-C(10)1.28(2)
O(4)-C(4)1.260(6)C(8)-C(9)1.365(7)
O(5)-C(11)1.252(6)C(8)-C(10C)1.550(5)
O(6)-C(11)1.261(5)C(9)-C(10)1.378(7)
C(4)-C(5)1.508(6)O(5)-C(15)1.32(2)
C(5)-C(6)1.386(7)O(6)-C(17)1.28(2)
C(5)-C(10)1.391(7)C(15)-C(16)1.393(7)
C(6)-C(7)1.393(7)C(15)-C(17B)1.551(5)
C(7)-C(8)1.392(6)C(16)-C(17)1.376(7)
C(8)-C(9)1.391(7)
C(9)-C(10)1.389(7)
C(8)-C(11)1.497(6)
Table 2. The parameters of porous structure of compound 1.
Table 2. The parameters of porous structure of compound 1.
ActivationSpecific Surface Area/m2·g−1Vpore/cm3·g−1Vads(CO2) a/cm3(STP)·g−1
LangmuirBETDFTTotal aDFT
2 h237.3111.261.60.08590.060478.9
6 h145.391.386.10.07270.044034.3
a at P/P0 = 0.95.
Table 3. Crystal data and structure refinement details for 1–4.
Table 3. Crystal data and structure refinement details for 1–4.
Compound12∙2DMF3∙0.6DMF4
FormulaC42H54La2N4O16C48H68La2N6O22C35.80H40.20Cl2La2N4.60 O16.60C54H82La2N4O16
Formula weight1148.711358.901149.241321.06
Crystal systemTriclinicTriclinicMonoclinicTriclinic
Space groupP-1P-1P21/nP-1
a, Å10.1266(10)10.6159(4)12.4139(5)10.471(2)
b, Å10.3128(10)11.4493(4)12.9185(5)13.1311(17)
c, Å12.3495(12)12.4139(4)14.7089(6)13.8254(18)
α, deg81.021(3)90.5437(13)90112.436(12)
β, deg75.024(3)108.8847(13)103.7170(11)92.458(16)
γ, deg68.767(3)92.3540(13)90108.076(16)
V, A31158.5(2)1426.01(9)2291.57(16)1642.0(5)
Z1121
dcalc, g/cm31.6461.5821.6661.336
θ range, °2.48–28.742.03–35.632.46–30.002.98–26.60
Crystal size, mm0.20 × 0.05 × 0.050.38 × 0.33 × 0.080.20 × 0.10 × 0.060.25 × 0.14 × 0.05
μ, mm−11.8921.5592.0271.344
Reflnscollected/unique18,483/598628,659/13,07931,260/668234,062/20,437
Unique reflns [I > 2σ(I)]568311,716529810,195
Rint0.01930.02420.05380.0755
S(F2)1.0571.0561.0621.048
R1, wR2 [I > 2σ(I)]0.0198, 0.04640.0334, 0.07550.0489, 0.11440.0731, 0.1726
R1, wR2 (all data)0.0220, 0.04710.0399, 0.07800.0696, 0.12130.1361, 0.1916
Δρmax/Δρmin, e/Å31.35/−0.611.84/−1.212.16/−1.491.72/−1.23
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.; 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. https://doi.org/10.3390/molecules26092486

AMA Style

Trofimova OY, Maleeva AV, Ershova IV, Cherkasov AV, Fukin GK, Aysin RR, Kovalenko KA, Piskunov AV. Heteroleptic LaIII Anilate/Dicarboxylate Based Neutral 3D-Coordination Polymers. Molecules. 2021; 26(9):2486. https://doi.org/10.3390/molecules26092486

Chicago/Turabian Style

Trofimova, Olesya Y., Arina V. Maleeva, Irina V. Ershova, Anton V. Cherkasov, Georgy K. Fukin, Rinat R. Aysin, Konstantin A. Kovalenko, and Alexandr V. Piskunov. 2021. "Heteroleptic LaIII Anilate/Dicarboxylate Based Neutral 3D-Coordination Polymers" Molecules 26, no. 9: 2486. https://doi.org/10.3390/molecules26092486

Article Metrics

Back to TopTop