Next Article in Journal
Synthesis and Structural Characterization of Layered Ni+1/+2 Oxides Obtained by Topotactic Oxygen Release on Nd2−xSrxNiO4−δ Single Crystals
Next Article in Special Issue
Dinuclear Molybdenum(VI) Complexes Based on Flexible Succinyl and Adipoyl Dihydrazones
Previous Article in Journal
Optoelectronic Device Modeling and Simulation of Selenium-Based Solar Cells under LED Illumination
Previous Article in Special Issue
Sulfonato Complex Formation Rather than Sulfonate Binding in the Extraction of Base Metals with 2,2′-Biimidazole: Extraction and Complexation Studies
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

New Polynuclear Coordination Compounds Based on 2–(Carboxyphenyl)iminodiacetate Anion: Synthesis and X-rays Crystal Structures

by
Sebastián Martínez
1,
Carlos Kremer
1,
Javier González-Platas
2 and
Carolina Mendoza
1,*
1
Área Química Inorgánica, Departamento Estrella Campos, Facultad de Química, Universidad de la República, Montevideo 11800, Uruguay
2
Departamento de Física, Instituto Universitario de Estudios Avanzados en Física Atómica, Molecular y Fotónica (IUDEA), MALTA Consolider Team, Servicio de Difracción de Rayos X, Universidad de La Laguna, 38200 Tenerife, Spain
*
Author to whom correspondence should be addressed.
Crystals 2023, 13(12), 1669; https://doi.org/10.3390/cryst13121669
Submission received: 31 October 2023 / Revised: 29 November 2023 / Accepted: 29 November 2023 / Published: 9 December 2023
(This article belongs to the Special Issue The Polyhedral Face of Coordination Chemistry)

Abstract

:
In the present work, novel polymeric copper(II) coordination compounds, namely [Cu2(cpida)(H2O)4][Cu(cpida)]·3H2O (1) (cpida3− = 2-(carboxyphenyl)iminodiacetate anion) and Na[Cu(cpida)] (3), were synthesized and characterized using infrared spectroscopy, thermogravimetric analysis, elemental analysis, and single-crystal X-ray diffraction. Compound 1 was obtained by slowly evaporating an aqueous solution of H3cpida, copper(II) sulfate, and NaOH at room temperature. The structural characterization revealed that 1 is an ionic entity formed by the [Cu(cpida)] anion and the [Cu2(cpida)(H2O)4]+ cation, both of polymeric 1D structure. Compound 3 was prepared under similar conditions from copper perchlorate and crystallized via acetone diffusion. It is a coordination polymer formed by the [Cu(cpida)] units, and the sodium cation is present in the structure, counterbalancing the anion charge. Depending on the crystallization conditions, it was possible to obtain other solvation forms of these structures. Starting from the conditions of compound 1, via the diffusion of ethanol, [Cu2(cpida)(H2O)4][Cu(cpida)]·H2O·1/2EtOH (2) was precipitated, while a hydrate form of compound 3 with the formula Na[Cu(cpida)]·2H2O (4) was obtained via methanol diffusion.

1. Introduction

Over the past few decades, there has been substantial development in the research on polynuclear coordination compounds and coordination polymers, which is particularly fueled by their potential applications as materials for gas storage, catalysis, environmental remediation, drug delivery, and in magnetochemistry or semiochemistry [1,2,3,4,5,6,7,8,9,10,11,12,13].
According to the IUPAC recommendations, coordination polymers are “coordination compounds with repeating coordination entities extending in 1, 2, or 3 dimensions” [14,15]. In the structures of these compounds, there are inorganic nodes (metal-containing clusters or metal ions) connected by organic divergent molecules (called linkers or connectors) [1,16,17]. Rigid ligands with a high degree of directionality were originally used as linkers in order to obtain specific crystal motifs [18]. However, the employment of more flexible ligands in the development of coordination polymers has lately been explored by numerous research groups [19,20,21,22,23,24,25,26,27,28,29,30,31,32,33]. The flexibility arises from the rotation of single bonds in carbon chains [22], allowing for various conformations that can result in distinct final structures [24].
The employment of flexible ligands as linkers in coordination polymers has encountered some hurdles when compared with the use of less flexible or rigid connectors [24]. This is due to the fact that the process of self-assembling is dependent on the reaction conditions, particularly the solution pH value, the type of metal cation, the concentrations of the different reactants, and the solvent employed. What is more, flexible polydentate ligands are also able to form mononuclear coordination compounds in competitive processes or adopt different conformations that are difficult to control.
Flexible molecules containing carboxylic acid groups are often chosen for the assembly of polynuclear coordination compounds. This preference is supported by the fact that this type of ligand can coordinate in different ways due to the possibility of fully or partially deprotonating the carboxylate groups, resulting in different topologies of the final structure [34]. These flexible molecules can act as bridging bidentate or tridentate ligands or chelating agents and even coordinate in a unidentate mode [35]. They are also good hydrogen bond acceptors capable of assembling supramolecular structures [36].
For several years now, we have focused our work on the development of polynuclear coordination compounds using polytopic flexible ligands. When working with aqueous solutions at mild acidic pH values, we reported the obtention of four capsule-like trinuclear copper(II) complexes assembled with nitrilotripropionic acid (H3ntp) [37,38] and three new metal polynuclear compounds (containing nickel(II) or copper (II)) with N-benzyliminodipropionic acid (H2bzlidp) [39]. In the case of the ntp complexes, the anion of copper salt had a significant influence on the obtained structure, defining the connectivity between the capsules [37]. When bzlidp2− was employed as a ligand, it was possible to obtain Cu(II) complexes whose structures were significantly different (dinuclear complex vs. 2D layered structure) depending on the starting conditions [39].
We then turned our attention to the tricarboxylic acid compound 2-(carboxyphenyl)iminodiacetic acid (H3cpida, Figure 1) as a connector of the metal ions. This is a multidentate ligand with the ability to form five-membered chelate rings. It presents flexible iminodiacetate groups and an aromatic ring that can initiate π–π interactions to stabilize the crystal packing.
This ligand has already been employed in the preparation of mononuclear [40,41,42,43] and trinuclear coordination compounds [44] and high-nuclearity metal aggregates [45,46,47,48,49]. However, the employment of this tricarboxylic molecule to connect metal centers in a 1D structure has been scarcely studied [44,50,51]. Bandyopadhyay and collaborators reported two polymeric copper (II) compounds, {[Cu(Hcpida)]·H2O}n and {[Cu(Hcpida)(H2O)]·H2O}n [50]. In the first compound, the copper centers are pentacoordinated, showing a square pyramid geometry with the Hcpida2− anion bridging the metal ions through the carboxylate groups attached to the phenyl ring. In this complex, one of the carboxylate groups remains protonated, stabilizing the chains through hydrogen bonds. In the second compound, the Cu(II) center is hexacoordinated, presenting a distorted octahedral geometry. The protonated carboxylate group is connected to the coordinated water molecule of a [Cu(Hcpida)(H2O)] unit of an adjacent chain through a hydrogen bond. In this way, cylindrical channels are formed in the structure that hosts the water crystallization molecule. Yong and coworkers described a Cd(II) coordination polymer, [Cd3(cpida)2(H2O)]·H2O, which contains three crystallographically independent Cd(II) centers and forms infinite helical chains [51]. Heteroleptic Cu(II) chain-like structures can also be found in the literature. Ma and coworkers characterized a Cu(II) coordination polymer containing the cpida3− anion and 1,10-phenantroline (phen) with a {[Cu2(cpida)(phen)(NO3)]·2H2O}n formula. The complex contains two five-coordinated copper(II) centers with a square pyramid geometry. Two cpida3− anions bridge three copper ions, forming a trinuclear triangular cluster that, in turn, is connected by their vertices, forming a 1D structure [44]. Polymeric coordination compounds based on the meta and para isomers of the 2-(carboxyphenyl)iminodiacetate anion have also been reported [35,52,53,54].
We decided to further explore the potential of H3cpida to form polynuclear systems, in conditions analogous to those employed in the preparation of the H2bzlidp or H3ntp compounds mentioned above. Copper(II) was used as a metal ion and we report the preparation of four new copper(II) polynuclear compounds, [Cu2(cpida)(H2O)4][Cu(cpida)]·3H2O (1), [Cu2(cpida)(H2O)4][Cu(cpida)]·H2O·1/2EtOH (2), Na[Cu(cpida)] (3), and Na[Cu(cpida)]·2H2O (4), obtained under different crystallization conditions.

2. Materials and Methods

2.1. General Information

All common laboratory chemicals were reagent grade, acquired from commercial suppliers, and employed without additional purification. Elemental analysis (C, H, N, S) was carried out on a Flash 2000 instrument (Thermo Fisher Scientific, USA). The infrared spectra, as KBr pellets, were obtained on an FTIR Shimadzu IR-Prestige-21 spectrophotometer (Shimadzu, Kyoto, Japan) in the 4000–400 cm−1 range. Thermogravimetric analyses (TGAs) were executed in a Shimadzu TGA-50 instrument with a TA 50 l interface (Shimadzu, Kyoto, Japan), using platinum cells. TGAs were recorded under a nitrogen stream (50 mL min−1) at 1.0 °C min−1 up to 300 °C and 30 °C min−1 from 300 to 700 °C.
CAUTION! Perchlorate salts of metal coordination compounds containing organic ligands have the potential to be explosive. Only small quantities of material should be prepared and it should be handled with caution.

2.2. Synthesis of the Complexes

2.2.1. Synthesis of [Cu2(cpida)(H2O)4][Cu(cpida)]·3H2O (1)

H3cpida (0.167 g, 0.66 mmol) was dissolved in distilled water (12 mL) and copper sulfate (0.159 g, 1.00 mmol) was added. The solution was magnetically stirred at room temperature and adjusted to pH 4.1 with 5 mol L–1 NaOH solution. The blue solution was filtered and the solvent was slowly evaporated at room temperature. After four months, 1 precipitated as green crystals. The crystals were filtered through paper and washed with cold distilled water (0 °C). The solid material was dried in open air, and 0.084 g of the product was obtained. The yield was 31% (based on copper(II)). Anal. Calc. for C22H30Cu3N2O19; C, 32.34; H, 3.70; N, 3.43. Found: C, 31.90; H, 3.45; N 3.39%. Selected IR data (νmax/cm−1): 3447, 3178, 2968 (sh), 2929 (sh), 1599, 1549, 1481, 1454, 1410, 1306, 1259, 1175, 1117, 1084, 1032, 1001, 990, 957, 932, 880, 812, 763, 723, 710, 646,594, 527, 446. TGA (30–700 °C, N2), % weight loss: 11.29 (left limit 19 °C–right limit 63 °C), 4.95 (left limit 63 °C–right limit 139 °C), 38.95 (left limit 170 °C –right limit 328 °C), 7.26% (left limit 328 °C–right limit 700 °C).

2.2.2. Synthesis of [Cu2(cpida)(H2O)4][Cu(cpida)]·H2O·1/2EtOH (2)

H3cpida (0.253 g, 1.0 mmol) was dissolved in distilled water (6 mL) and the pH was set to 7 through addition of 5 mol L–1 NaOH. This solution was added to a 2 mL aqueous solution of copper trifluoromethanesulfonate (0.543 g, 1.5 mmol). The solution was stirred at room temperature and its pH adjusted to 5.2 with aqueous 5 mol L–1 NaOH solution. This solution was subsequently divided into two portions of 4 mL each, which were set to crystalize either by diffusion of ethanol or methanol vapors. After 21 d, green crystals were obtained from the first (ethanol vapors). The crystals were filtered through paper and washed with cold distilled water (0 °C). After allowing the crystals to dry in open air, 0.070 g of 2 was obtained. The yield was 35% (based on copper(II)). Anal. Calc. for C46H58Cu6N4O35: C, 34.35; H, 3.64; N, 3.48. Found: C, 33.54; H, 3.61; N 3.34%. Selected IR data (νmax/cm−1): 3447, 3145 (sh), 2972, 2928, 2855, 1609, 1564, 1547, 1479, 1454, 1408, 1373, 1304, 1258, 1173, 1121, 1086, 1040, 999, 957, 934, 876, 808,766, 721, 712, 646, 590, 525.
Blue needle-shaped crystals of a different compositions were obtained from the diffusion of methanol (see below).

2.2.3. Synthesis of Na[Cu(cpida)] (3)

H3cpida (0.253 g, 1.0 mmol) was dissolved in 6 mL of distilled water and the pH value of the resulting solution was set to 7 by adding NaOH (5 mol L–1 solution). This ligand solution was added to a 2 mL aqueous solution of copper perchlorate hexahydrate (0.556 g, 1.5 mmol). The solution was stirred at room temperature and its pH was set to 5.2, again with 5 mol L–1 NaOH solution. Blue needle-shaped crystals were obtained after 60 days, produced from 2 mL of this solution via diffusion of acetone vapors. These crystals were filtered, washed with cold distilled water (0 °C), and allowed to dry in open air. Following this procedure, 0.034 g of the product was obtained. The yield was 40% (based on H3cpida). Anal. Calc. for C11H8CuNNaO6; C, 39.24; H, 2.39; N, 4.16. Found: % C, 38.82; H, 2.42; N 4.06%. Selected IR data (νmax/cm−1): 3441, 2979, 2939, 1674, 1638, 1616, 1483, 1450, 1396, 1385, 1339, 1219, 1171, 1152, 1103, 1061, 1045, 961, 941, 908, 881, 853, 777, 752, 669, 617, 575, 550, 523, 442.

2.2.4. Synthesis of Na[Cu(cpida)]·2H2O (4)

H3cpida (0.253 g, 1.0 mmol) was dissolved in 6 mL of distilled water and the pH was set to 7 with the addition of NaOH (5 mol L–1 solution), drop by drop. This solution was added to 2 mL of aqueous copper perchlorate hexahydrate solution (0.556 g, 1.5 mmol). The mixture was magnetically stirred at room temperature and the pH was set to 5.2 with aqueous NaOH solution. Blue needle-shaped crystals of compound 4 were obtained after 50 days, produced from 2 mL of the resulting solution via diffusion of methanol vapors. The crystals were separated by filtration, washed with cold water (0 °C), and dried in open air. Following this procedure, 0.035 g of the product was obtained. The yield was 38% (based on H3cpida) Anal. Calc. for C11H12CuNNaO8, C, 35.44; H, 3.25; N, 3.76. Found: C, 35.05; H, 3.12; N, 3.54%. Selected IR data (νmax/cm−1): 3423, 2970, 2932, 1657, 1634, 1481, 1447, 1402, 1369, 1346, 1317, 1228, 1182, 1163, 1148, 1099, 1063, 1049, 982, 959, 941, 910, 880, 853, 766, 752, 708, 671, 652, 617, 578, 552, 529, 492, 447. TGA (30–700 °C, N2) % weight loss: 8.00 (left limit 19 °C–right limit 120 °C), 34.66 (left limit 100 °C–right limit 305 °C), 20.96 (left limit 305 °C–right limit 700 °C).
Compound 4 could also be obtained from copper trifluoromethanesulfonate, in a similar manner as compound 2 but using methanol for the diffusion (0.0315 g).

2.3. X-ray Crystallography

The X-ray diffraction data for all compounds were collected with a Rigaku SuperNOVA diffractometer with microfocus X-ray using Cu/Mo Kα radiation (λ = 1.54184/0.71073 Å). CrysAlisPro 1.171.39.46 [55] software was employed to collect, index, scale, and apply analytical absorption correction based on the faces of the crystal. The SHELXT program was used to solve the structure with a dual-space algorithm [56]. Fourier recycling and least-squares refinement were used for the model completion with SHELXL-2019 [57]. All non-hydrogen atoms were refined anisotropically, and all hydrogen atoms were placed in geometrically suitable positions and refined through riding with the isotropic thermal parameter related to the equivalent isotropic thermal parameter of the parent atom. The hydrogen atoms were geometrically positioned with C-H = 0.93 Å and Uiso(H) = 1.2 Ueq(C). The geometrical analysis of interactions in the structure was performed with the Olex2 [58] program (version 1.5).
The crystal data, collection procedures, and refinements results are summarized in Table 1.

3. Results and Discussion

3.1. Synthesis of the Complexes

The coordination compounds were prepared by directly mixing aqueous solutions of the corresponding copper salts and H3cpida, setting the pH value in the 4.0–5.2 range, and using a 1.5 copper-to-ligand ratio (Scheme 1).
The crystals were characterized via infrared spectroscopy and elemental analysis. IR spectra of compounds 1, 2, and 4 exhibit a broad band in the 3200–3600 cm−1 region. The presence of this band is indicative of the existence of water molecules in the crystal structures. The small absorption peaks that appeared in the 3000–2850 cm−1 region for the four complexes correspond to the νC–H vibration modes of the –CH2– groups in the alkyl chains on the cpida3− anion (Figures S1–S4 Supplementary Material). The absence of the distinctive band of νas(COOH) at ca. 1700 cm−1 indicates that the carboxylate groups of the ligand are fully deprotonated in these compounds. For complexes 1 and 2, the νas(COO) vibrations appeared at ca. 1600 cm−1, whereas those of νs(COO) occurred at ca. 1410 cm−1. In the case of compounds 3 and 4, the bands around 1635 cm−1 can be assigned to the asymmetric stretching of the COO of the complexes while those that appeared in the proximity of 1400 cm−1 correspond to νs(COO). In all cases, the elemental analysis is consistent with the proposed empirical formula.
The TGA curve for compound 1 (Figure S5, Supplementary Material) shows, below 140 °C, a dehydration process that occurs in two steps, with a weight loss of 16.2%. The dehydration process is consistent with the loss of seven water molecules, comprising both lattice water molecules and coordinated ones. This is followed by a 46% weight loss step that involves the decomposition of the organic ligand. This is similar to what has previously been observed in the TG analysis of complexes containing the cpida ligand [43,44,50,51,53]. In the case of compound 4, the TGA curve (Figure S6, Supplementary Material) shows a weight loss of 8% below 120 °C, consistent with a two-molecule dehydration process, and a two-step decomposition process with a weight loss of circa 56%, corresponding to the decomposition of the organic ligand. Single crystals for the complexes were obtained through slow evaporation of the solvent (1), and slow diffusion of ethanol (2), acetone (3), and methanol (4), at room temperature. Crystallographic data of compounds 1 to 4 are summarized in Table 1.

3.2. Crystal Structure of [Cu2(cpida)(H2O)4][Cu(cpida)]·3H2O (1) and of [[Cu2(cpida)(H2O)4][Cu(cpida)]·H2O·1/2EtOH] (2)

Compound 1 crystallizes in the monoclinic P 21/n space group (Table 1). Selected bond lengths and angles for 1 are presented in Table 2 and in Table S1 (Supporting Information). The asymmetric unit of [Cu2(cpida)(H2O)4][Cu(cpida)] is shown in Figure 2a. Compound 1 is a salt formed by the [Cu(cpida)] anion and the [Cu2(cpida)(H2O)4]+ cation, both of them being polymeric, in a type of structure that is quite common with this kind of iminocarboxylic ligand [39,50]. The structure also presents three lattice water molecules. The anion is formed by [Cu(cpida)] units, connected by carboxylate oxygen atoms. Each unit contains a Cu(II) ion coordinated to a cpida3− anion, which acts as a tetradentate ligand. In order to assess the coordination geometry of Cu2, which shows five ligated atoms, the τ factor was determined. This is a geometrical parameter proposed by Addison and coworkers [59] to classify five coordinated structures in square pyramidal (τ factor = 0) or trigonal bipyramids (τ factor = 1). According to this parameter, the coordination polyhedron of the Cu2 atom can be depicted as a slightly distorted square pyramid (τ factor = 0.33), the donor atoms being the nitrogen atom N2, and three oxygen atoms of each carboxylate of the cpida3− anion (O8, O9, and O12, Figure 2b). The coordination sphere is completed by the O-carboxylate of a neighboring [Cu(cpida)] unit. Therefore, two of the carboxylate groups are coordinating in a monodentate mode, while the third is acting as bridging bidentate, in a syn-anti conformation and with a Cu–Cu separation of 5.25(7) Å. Cu(II) ida or substituted ida complexes have shown this kind of connection [44,60,61,62] as opposed to the one previously reported [50] where the bridging carboxylate group was the one attached to the phenyl ring.
The [Cu2(cpida)(H2O)4]+ cation is dinuclear and it is formed by [Cu(cpida)] units connected by carboxylate oxygen atoms. The coordination around Cu1 is also a distorted square pyramid (τ factor = 0.278) [59], with the donor atoms being three oxygen atoms, one of each carboxylate of the cpida3− ligand (O2, O3, and O6), and the nitrogen atom N1 (Figure 2c). The fifth coordination position is occupied by the O1 atom of a neighboring acetic group that is acting as a bridging bidentate group, in a syn-anti conformation and with a Cu1–Cu1 separation of 5.17(7) Å. In this way, a 1D chain of [Cu(cpida)] is formed, which is similar to the anionic chain in the structure. The cation presents a second Cu atom (Cu3), which is pentacoordinated, again with a very slightly distorted square pyramid (τ factor = 0.002) [59]. The donor atoms are the oxygen atoms of four water molecules (O1W, O2W, O3W, and O4W) and the O5 atom of a carboxylate group. This carboxylate group, which is directly attached to the phenyl ring, is bridging Cu1 and Cu3 in a syn-anti conformation. The equatorial positions of the square pyramid are occupied by the O5 atom and three of the coordinated water molecules (O2W, O3W, and O4W), while the fourth water molecule (O1W) is placed at the axial position. The crystal packing is reinforced by hydrogen bonds between the lattice water molecules and the carboxylate oxygen atoms with an average donor acceptor distance of d(D·· A) = 2.74(8) Å and an average D-H· ·A angle of 161.8° (Table S3, Supporting Information).
Compound 2 was obtained under similar conditions as those used for compound 1, save for the copper source and diffusion vapors. Selected bond lengths and angles for compound 2 are showed in Table 2 and in Table S1 (Supporting Information). The compound crystallizes in the monoclinic P 21/n space group (Table 1) and its structure is similar to that of compound 1, with the difference that just one lattice water molecule and half a molecule of ethanol are present per asymmetric unit (Figure 3a). Compound 2 is also composed of the [Cu2(cpida)(H2O)4]+ cation and the [Cu(cpida)] anion, both presenting a chain-like structure. In the [Cu2(cpida)(H2O)4]+ cation, the coordination polyhedron around Cu1 is also a distorted square pyramid (τ factor = 0.281) [59], with the donor atoms being three oxygen atoms of each carboxylate group of the cpida3− anion (O2, O3, and O6, Figure 3a) and the nitrogen atom N1. The fifth coordination position is occupied by the O1 atom of a neighboring acetic group, which is acting as a bridging bidentate group in a syn-anti conformation, and with a Cu1–Cu1 separation of 5.17(7) Å. The Cu atom (Cu3) is pentacoordinated, and again with a slightly distorted square pyramid geometry (τ factor = 0.01) [59]. The donor atoms are the oxygen atoms of four water molecules (O1W, O2W, O3W, and O4W) and O5 of the carboxylate group attached to the phenyl ring. This carboxylate group is bridging Cu1 and Cu3 in a syn-anti conformation. Three water molecules (O1W, O2W, and O4W) and the O5 atom are occupying the equatorial positions of the square pyramid, while the fourth coordinated water molecule (O3W) is placed at the axial position.
In the [Cu(cpida)] anion, the coordination geometry around the Cu2 atom can be described as a slightly distorted square pyramid (τ factor = 0.36) [59], with the donor atoms being the nitrogen atom N2 and three oxygen atoms, one of each carboxylate group of the cpida3− anion (O8, O9, and O12, Figure 3a). The coordination polyhedron is completed by the O-carboxylate of a neighboring [Cu(cpida)] unit. One of the acetic carboxylate groups is acting as a bridging bidentate, in a syn-anti conformation and with a Cu–Cu separation of 5.25(7) Å.
The crystal packing is reinforced by hydrogen bonds between the lattice water molecules, the ethanol molecules, and the carboxylate oxygen atoms, with an average donor acceptor distance of d(D· ·A) = 2.73(5) Å and an average D-H· ·A angle of 162.2° (Table S4, Supporting Information). An unusual C(4)-H(4B)· ·O(13)#1 hydrogen bond between the ethanol O(13) atom and the CH2 group of the cationic chain is also present (Table S4).
Given that the unit cells of 1 and 2 showed practically the same geometrical parameters and the same symmetry, we performed an analysis to further characterize these similarities. Using the COMPSTRU feature available on the Bilbao Crystallographic Server (https://www.cryst.ehu.es/ accessed on 14 April 2023), we checked the relationship between both structures, after removing the corresponding solvent molecules. It was found that there is no pair of atoms transformable with a one-angstrom tolerance. Mercury’s structure overlay tool (Mercury version 2022.3.0) [63] was also employed, with us observing that when increasing the number of pairs of atoms to overlap, the RMSD between the chosen molecules increased more and more.

3.3. Crystal structure of Na[Cu(cpida)] (3)

Compound 3 crystallizes in an orthorhombic Pbca space group (Table 1). Selected bond lengths and angles for 3 can be found in Table 3 and Table S2 (Supporting Information). The asymmetric unit (Figure 4a) is composed of the [Cu(cpida)] anion with the cpida3− ligand coordinated to the Cu cation in a tridentate manner. The coordination sphere of the copper center is square planar, with the donor atoms being N1, O5, O3, and the O1#5 atom of a neighboring [Cu(cpida)] unit, with an average Cu-O distance of 1.92(4) Å and a Cu-N distance of 2.046(5) Å. The O1 atom is located in an axial position, at a longer Cu1-O1 distance. Therefore, the [Cu(cpida)] fragments form a chain-like structure, in a zig-zag motif (Figure 4b). In the structure, there are also sodium cations, which are connected to the carboxylate oxygen atoms O5 and O1 and to the O(3)#6, O(3)#7, O(6)#8, O(4)#6, and O(2)#5 atoms, connecting the [Cu(cpida)] chains (Figure 4c). The packing of 3 viewed along the b axis is depicted in Figure 4d.

3.4. Crystal Structure of Na[Cu(cpida)]·2H2O (4)

Compound 4 crystallizes in the orthorhombic Iba2 space group (Table 1). Selected bond lengths and angles for 4 are displayed in Table 3 and Table S2 (Supporting Information). The composition of the asymmetric unit (Figure 5a) is like that of 3. It contains the [Cu(cpida)] anion with the cpida3− ligand coordinated to the Cu cation in a tridentate fashion. The coordination sphere of the copper center is square planar, with the donor atoms being N1, O5, O3, and the O1#9 atom of a neighboring [Cu(cpida)] unit, with a Cu-N distance of 2.054(3) Å and an average Cu-O distance of 1.93(2) Å. The O1 atom is located in an axial position, at a longer Cu1-O1 distance. Again, the [Cu(cpida)] fragments form a chain-like structure, in a zig-zag motif (Figure 5b). In the structure, there are also sodium cations that are connected to the carboxylate oxygen atoms O5, O1, O(3)#10, O(3)#11, O(6)#12, O(4)#10, and O(2)#9, resulting in the overall stabilization of the [Cu(cpida)] chains (Figure 5c). Water crystallization molecules are present in the unit cell but, unfortunately, it was impossible to model it correctly and, therefore, a squeeze method was applied in this case. The analysis of the solvent mask we calculated gave 136 e in a volume of 446 Å3 in a void per unit cell. This is consistent with the presence of two water molecules per asymmetric unit. The total amount of solvent calculated through this procedure (solvent mask, squeeze) also reasonably matches the elemental analysis and the 8.0% mass loss observed in TGA up to 120 °C (Figure S6), with a calculated mass of 9.6% (two H2O molecules per Na[Cu(cpida)]·2H2O). The packing of 4 is depicted in Figure 5d.

4. Conclusions

In this work, we have reported the structure of four new Cu(II) polymeric complexes derived from the 2-(carboxyphenyl)iminodiacetic acid, H3cpida, obtained under different crystal growing conditions.
Compounds [Cu2(cpida)(H2O)4][Cu(cpida)]·3H2O (1) and [Cu2(cpida)(H2O)4][Cu(cpida)]·H2O·1/2EtOH (2), obtained through the evaporation of the solvent and ethanol vapor diffusion, respectively, are ionic, formed by the [Cu(cpida)] anion and the [Cu2(cpida)(H2O)4]+ cation, both of them 1D coordination polymers. The structure of the anion is based on [Cu(cpida)] units connected by an acetic oxygen atom that is bridging the Cu centers in a chain-like structure. The [Cu2(cpida)(H2O)4]+ cation is dinuclear and it is formed by [Cu(cpida)] units connected by one of the acetic carboxylate groups, while the phenyl carboxylate group is bridging a second copper (II) ion whose coordination sphere is completed with water molecules. Solvent molecules are present in the unit cells of both compounds.
Through diffusion of acetone or methanol vapors, compounds 3, Na[Cu(cpida)], and 4, Na[Cu(cpida)]·2H2O, are crystallized. They are also polymeric, with one of the acetic carboxylate groups of the ligand bridging the Cu(II) ions, forming an anionic chain-like structure. The sodium cations are balancing the negative charge, interacting with the oxygen atoms of different carboxylate groups, and connecting the [Cu(cpida)]n chains.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/cryst13121669/s1, Figure S1: FT-IR spectrum of compound 1; Figure S2: FT-IR spectrum of compound 2; Figure S3: FT-IR spectrum of compound 3; Figure S4: FT-IR spectrum of compound 4; Figure S5: TGA diagram of compound 1; Figure S6: TGA diagram of compound 4; Table S1. Selected bond angles (°) for the coordination center in compounds 1 and 2; Table S2. Selected bond angles (°) for the coordination center in compounds 3 and 4. Table S3: Hydrogen bonds of 1; Table S4: Hydrogen bonds for 2.

Author Contributions

Writing—original draft preparation, C.M; writing—review and editing, C.M., S.M., J.G.-P., and C.K.; synthesis and characterization, S.M. and C.M.; diffraction studies, J.G.-P.; conceptualization, C.K and C.M. All authors have read and agreed to the published version of the manuscript.

Funding

This work was partially supported by ANII (Agencia Nacional de Investigación e Innovación, Uruguay) via Project FCE_2_2011_1_N° 6638 and CSIC (Comisión Sectorial de Investigación Científica, Uruguay) through Programa de Apoyo a Grupos de Investigación.

Data Availability Statement

Data are contained within the article and supplementary materials.

Acknowledgments

We are grateful for the support of PEDECIBA (Programa para el Desarrollo de las Ciencias Básicas, Uruguay). J.G.-P. is thankful for the support of MCIN/AEI/10.13039/5011000011033 through the project PID2019-106383GB-C44. C.M. wants to thank Natalia Alvarez (Área Química Inorgánica, DEC, Facultad de Química, Udelar), for the valuable discussion of the results.

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. Ma, L.; Abney, C.; Lin, W. Enantioselective catalysis with homochiral metal–organic frameworks. Chem. Soc. Rev. 2009, 38, 1248–1256. [Google Scholar] [CrossRef] [PubMed]
  2. Chen, W.; Wu, C. Synthesis, functionalization, and applications of metal–organic frameworks in biomedicine. Dalton Trans. 2018, 47, 2114–2133. [Google Scholar] [CrossRef] [PubMed]
  3. Baruah, J.B. Coordination polymers in adsorptive remediation of environmental contaminants. Coord. Chem. Rev. 2022, 470, 214694. [Google Scholar] [CrossRef]
  4. Nguyen, N.T.T.; Nguyen, T.T.T.; Nguyen, D.T.C.; Van Tran, T. Functionalization strategies of metal-organic frameworks for biomedical applications and treatment of emerging pollutants: A review. Sci. Total Environ. 2024, 906, 167295. [Google Scholar] [CrossRef]
  5. Dutta, M.; Bora, J.; Chetia, B. Overview on recent advances of magnetic metal–organic framework (MMOF) composites in removal of heavy metals from aqueous system. Environ. Sci. Pollut. Res. 2022, 30, 13867–13908. [Google Scholar] [CrossRef]
  6. Li, J.-R.; Kuppler, R.J.; Zhou, H.-C. Selective gas adsorption and separation in metal–organic frameworks. Chem. Soc. Rev. 2009, 38, 1477–1504. [Google Scholar] [CrossRef]
  7. Farrusseng, D.; Aguado, S.; Pinel, C. Metal–Organic Frameworks: Opportunities for Catalysis. Angew. Chem. Int. Ed. 2009, 48, 7502–7513. [Google Scholar] [CrossRef]
  8. Ma, Z.; Moulton, B. Recent advances of discrete coordination complexes and coordination polymers in drug delivery. Coord. Chem. Rev. 2011, 255, 1623–1641. [Google Scholar] [CrossRef]
  9. Kurmoo, M. Magnetic metal–organic frameworks. Chem. Soc. Rev. 2009, 38, 1353–1379. [Google Scholar] [CrossRef]
  10. Jiang, H.-L.; Xu, Q. Porous metal–organic frameworks as platforms for functional applications. Chem. Commun. 2011, 47, 3351–3370. [Google Scholar] [CrossRef]
  11. Murray, L.J.; Dincă, M.; Long, J.R. Hydrogen storage in metal–organic frameworks. Chem. Soc. Rev. 2009, 38, 1294–1314. [Google Scholar] [CrossRef] [PubMed]
  12. Loukopoulos, E.; Kostakis, G.E. Review: Recent advances of one-dimensional coordination polymers as catalysts. J. Coord. Chem. 2018, 71, 371–410. [Google Scholar] [CrossRef]
  13. Pettinari, C.; Marchetti, F.; Mosca, N.; Tosi, G.; Drozdov, A. Application of metal–Organic frameworks. Polym. Int. 2017, 66, 731–744. [Google Scholar] [CrossRef]
  14. Batten, S.R.; Champness, N.R.; Chen, X.-M.; 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]
  15. Batten, S.R.; Champness, N.R.; Chen, X.-M.; Garcia-Martinez, J.; Kitagawa, S.; Öhrström, L.; O’Keeffe, M.; Suh, M.P.; Reedijk, J. Coordination polymers, metal–organic frameworks and the need for terminology guidelines. CrystEngComm 2012, 14, 3001–3004. [Google Scholar] [CrossRef]
  16. Yusuf, V.F.; Malek, N.I.; Kailasa, S.K. Review on Metal–Organic Framework Classification, Synthetic Approaches, and Influencing Factors: Applications in Energy, Drug Delivery, and Wastewater Treatment. ACS Omega 2022, 7, 44507–44531. [Google Scholar] [CrossRef]
  17. Masoomi, M.Y.; Morsali, A. Applications of metal–organic coordination polymers as precursors for preparation of nano-materials. Coord. Chem. Rev. 2012, 256, 2921–2943. [Google Scholar] [CrossRef]
  18. Robson, R. Design and its limitations in the construction of bi- and poly-nuclear coordination complexes and coordination polymers (aka MOFs): A personal view. Dalton Trans. 2008, 38, 5113–5131. [Google Scholar] [CrossRef]
  19. Zhang, M.-L.; Zheng, Y.-J.; Ma, Z.-Z.; Ren, Y.-X.; Cao, J.; Wang, Z.-X.; Wang, J.-J. Zinc(II) and cadmium(II) complexes of long flexible bis(imidazole) and phenylenediacetate ligands, synthesis, structure, and luminescent property. Polyhedron 2018, 146, 180–186. [Google Scholar] [CrossRef]
  20. Khalaj, M.; Lalegani, A.; Akbari, J.; Ghazanfarpour-Darjani, M.; Lyczko, K.; Lipkowski, J. Synthesis and characterization of three new Cd(II) coordination polymers with bidentate flexible ligands: Formation of 3D and 1D structures. J. Mol. Struct. 2018, 1169, 31–38. [Google Scholar] [CrossRef]
  21. Chen, M.-L.; Zhou, Z.-H. Structural diversity of 1,3-propylenediaminetetraacetato metal complexes: From coordination monomers to coordination polymers and MOF materials. Inorg. Chim. Acta 2017, 458, 199–217. [Google Scholar] [CrossRef]
  22. Li, H.; Tian, H.; Guo, M.; He, F.Y.; Hu, C. Self-complementary self-assembly of 3D coordination framework involving d-block, s-block metal ions and flexible NTA3−ligand. Inorg. Chem. Commun. 2006, 9, 895–898. [Google Scholar] [CrossRef]
  23. Liu, T.; Lü, J.; Shi, L.; Guo, Z.; Cao, R. Conformation control of a flexible 1,4-phenylenediacetate ligand in coordination complexes: A rigidity-modulated strategy. CrystEngComm 2008, 11, 583–588. [Google Scholar] [CrossRef]
  24. Liu, T.-F.; Lü, J.; Cao, R. Coordination polymers based on flexible ditopic carboxylate or nitrogen-donor ligands. CrystEngComm 2010, 12, 660–670. [Google Scholar] [CrossRef]
  25. Mateescu, A.; Gabriel, C.; Raptis, R.G.; Baran, P.; Salifoglou, A. pH—Specific synthesis, spectroscopic, and structural characterization of an assembly of species between Co(II) and N,N-bis(phosphonomethyl)glycine. Gaining insight into metal-ion phosphonate interactions in aqueous Co(II)–organophosphonate systems. Inorg. Chim. Acta 2007, 360, 638–648. [Google Scholar] [CrossRef]
  26. Zhang, Q.-Z.; Lu, C.-Z.; Yang, W.-B.; Yu, Y.-Q. A novel three-dimensional framework formed by polymeric aqua (nitrilotriacetato)-lanthanum: [La(C6H6NO6)H2O]n. Inorg. Chem. Commun. 2004, 7, 277–279. [Google Scholar] [CrossRef]
  27. Carballo, R.; Covelo, B.; El Fallah, M.S.; Ribas, J.; Vázquez-López, E.M. Supramolecular Architectures and Magnetic Behavior of Coordination Polymers from Copper(II) Carboxylates and 1,2-Bis(4-pyridyl)ethane as a Flexible Bridging Ligand. Cryst. Growth Des. 2007, 7, 1069–1077. [Google Scholar] [CrossRef]
  28. Hawes, C.S.; Chilton, N.F.; Moubaraki, B.; Knowles, G.P.; Chaffee, A.L.; Murray, K.S.; Batten, S.R.; Turner, D.R. Coordination polymers from a highly flexible alkyldiamine-derived ligand: Structure, magnetism and gas adsorption studies. Dalton Trans. 2015, 44, 17494–17507. [Google Scholar] [CrossRef]
  29. Lai, L.-Y.; Liu, Z.; Han, G.-C.; Chen, Z. Synthesis, Crystal Structure and Properties of Three Metal Complexes Based on a Flexible Schiff Base Ligand. J. Clust. Sci. 2015, 26, 1845–1855. [Google Scholar] [CrossRef]
  30. Li, H.-Y.; Cao, L.-H.; Wei, Y.-L.; Xu, H.; Zang, S.-Q. Construction of a series of metal–organic frameworks based on novel flexible ligand 4-carboxy-1-(3,5-dicarboxy-benzyl)-pyridinium chloride and selective d-block metal ions: Crystal structures and photoluminescence. CrystEngComm 2015, 17, 6297–6307. [Google Scholar] [CrossRef]
  31. Karmakar, A.; Oliver, C.L.; Roy, S.; Öhrström, L. The synthesis, structure, topology and catalytic application of a novel cubane-based copper(ii) metal–organic framework derived from a flexible amido tripodal acid. Dalton Trans. 2015, 44, 10156–10165. [Google Scholar] [CrossRef]
  32. Yu, M.; Xie, L.; Liu, S.; Wang, C.; Cheng, H.; Ren, Y.; Su, Z. Photoluminescent metal-organic framework with hex topology constructed from infinite rod-shaped secondary building units and single e,e-trans-1,4-cyclohexanedicarboxylic dianion. Inorg. Chim. Acta 2007, 360, 3108–3112. [Google Scholar] [CrossRef]
  33. Xie, Y.; Bai, F.; Xing, Y.; Wang, Z.; Zhao, H.; Shi, Z. Synthesis and Crystal Structure of Novel Coordination Polymers with Nitrilotripropionic Acid. Z. Anorg. Allg. Chem. 2010, 636, 1585–1590. [Google Scholar] [CrossRef]
  34. Huang, R.-W.; Li, B.; Zhang, Y.-Q.; Zhao, Y.; Zang, S.-Q.; Xu, H. Divalent cobalt, zinc, and copper coordination polymers based on a new bifunctional ligand: Syntheses, crystal structures, and properties. Inorg. Chem. Commun. 2014, 39, 106–109. [Google Scholar] [CrossRef]
  35. Chai, X.; Zhang, H.; Zhang, S.; Cao, Y.; Chen, Y. The tunable coordination architectures of a flexible multicarboxylate N-(4-carboxyphenyl)iminodiacetic acid via different metal ions, pH values and auxiliary ligand. J. Solid State Chem. 2009, 182, 1889–1898. [Google Scholar] [CrossRef]
  36. Pan, Z.; Zheng, H.; Wang, T.; Song, Y.; Li, Y.; Guo, Z.; Batten, S.R. Hydrothermal Synthesis, Structures, and Physical Properties of Four New Flexible Multicarboxylate Ligands-Based Compounds. Inorg. Chem. 2008, 47, 9528–9536. [Google Scholar] [CrossRef]
  37. Braña, E.; Mendoza, C.; Vitoria, P.; González-Platas, J.; Domínguez, S.; Kremer, C. Trinuclear Cu(II) capsules self-assembled by nitrilotripropionate. Inorg. Chim. Acta 2014, 417, 192–200. [Google Scholar] [CrossRef]
  38. Martínez, S.; Veiga, N.; Torres, J.; Kremer, C.; Mendoza, C. Polynuclear complexes in solution: An experimental and theoretical study on the interaction of nitrilotripropionate anion with metal ions. Inorg. Chim. Acta 2018, 483, 53–60. [Google Scholar] [CrossRef]
  39. Braña, E.; Quiñone, D.; Martínez, S.; Grassi, J.; Carrera, I.; Torres, J.; González-Platas, J.; Seoane, G.; Kremer, C.; Mendoza, C. New polynuclear compounds based on N-benzyliminodipropionic acid: Solution studies, synthesis, and X-ray crystal structures. J. Coord. Chem. 2016, 69, 3650–3663. [Google Scholar] [CrossRef]
  40. Bera, M.; Musie, G.T.; Powell, D.R. Zinc(II) mediated cyclization and complexation of an unsymmetrical dicarboxyamine ligand: Synthesis, spectral and crystal structures characterizations. Inorg. Chem. Commun. 2008, 11, 293–299. [Google Scholar] [CrossRef]
  41. Palanisami, N.S.K.K.; Gopalakrishnan, T.; Moon, I.-S. A mixed Ni(II) ionic complex containing V-shaped water trimer: Synthesis, spectral, structural and thermal properties of {[Ni(2,2′-bpy)3][Ni(2-cpida)(2,2′ -bpy)]} (ClO4).3H2O. J. Chem. Sci. 2015, 127, 873–876. [Google Scholar] [CrossRef]
  42. Tomita, T.; Kyuno, E.; Tsuchiya, R. The Chromium(III) Complexes with Anthranilicdiacetic Acid. Bull. Chem. Soc. Jpn. 1969, 42, 947–951. [Google Scholar] [CrossRef]
  43. Chatterjee, C.; Singh, R.S.; Phulambrikar, A.; Das, S. Synthesis and spectral characterization of cobalt(III) complexes of N-(o-carboxyphenyl)iminodiacetic acid. J. Chem. Soc. Dalton Trans. 1988, 1988, 159–162. [Google Scholar] [CrossRef]
  44. Ma, J.; Jiang, F.; Zhou, K.; Chen, L.; Wu, M.; Hong, M. Effects of Temperature and Anion on the Copper(II) Complexes based on 2-(Carboxyphenyl)iminodiacetic Acid and 1,10-Phenanthroline. Z. Anorg. Und Allg. Chem. 2015, 641, 1998–2004. [Google Scholar] [CrossRef]
  45. Nallasamy, P.; Senthilkumar, K.; Mohan, G.; Moon, I.-S. Structural and luminescent properties of a tetranuclear cage-type cadmium(II) carboxylate cluster containing a V-shaped water trimer. J. Coord. Chem. 2016, 69, 1005–1013. [Google Scholar] [CrossRef]
  46. Murugesu, M.; Anson, C.E.; Powell, A.K. Engineering of ferrimagnetic Cu12-cluster arrays through supramolecular interactions. Chem. Commun. 2002, 10, 1054–1055. [Google Scholar] [CrossRef] [PubMed]
  47. Murugesu, M.; King, P.; Clérac, R.; Anson, C.E.; Powell, A.K. A novel nonanuclear CuIIcarboxylate-bridged cluster aggregate with an S = 7/2 ground spin state. Chem. Commun. 2004, 6, 740–741. [Google Scholar] [CrossRef]
  48. Murugesu, M.; Clérac, R.; Anson, C.E.; Powell, A.K. Polycopper(II) aggregates as building blocks for supramolecular magnetic structures. J. Phys. Chem. Solids 2004, 65, 667–676. [Google Scholar] [CrossRef]
  49. McCowan, C.S.; Groy, T.L.; Caudle, M.T. Synthesis, Structure, and Preparative Transamination of Tetrazinc Carbamato Complexes Having the Basic Zinc Carboxylate Structure. Inorg. Chem. 2002, 41, 1120–1127. [Google Scholar] [CrossRef]
  50. Bandyopadhyay, S.; Das, A.; Mukherjee, G.; Cantoni, A.; Bocelli, G.; Chaudhuri, S.; Ribas, J. Synthesis, X-ray crystal structures, physicochemical and magnetic studies on copper(II) N-(2-carboxyphenyl) iminodiacetate polymers. Polyhedron 2004, 23, 1081–1088. [Google Scholar] [CrossRef]
  51. Yong, G.-P.; Qiao, S.; Wang, Z.-Y.; Cui, Y. Five-, seven-, and eight-coordinate Cd(II) coordination polymers built by anthranilic acid derivatives: Synthesis, structures and photoluminescence. Inorg. Chim. Acta 2005, 358, 3905–3913. [Google Scholar] [CrossRef]
  52. Chu, Q.; Liu, G.-X.; Okamura, T.-A.; Huang, Y.-Q.; Sun, W.-Y.; Ueyama, N. Structure modulation of metal–organic frameworks via reaction pH: Self-assembly of a new carboxylate containing ligand N-(3-carboxyphenyl)iminodiacetic acid with cadmium(II) and cobalt(II) salts. Polyhedron 2008, 27, 812–820. [Google Scholar] [CrossRef]
  53. Yong, G.; Wang, Z.; Cui, Y. Synthesis, structural characterization and properties of copper(II) and zinc(II) coordination polymers with a new bridging chelating ligand. Eur. J. Inorg. Chem. 2004, 2004, 4317–4323. [Google Scholar] [CrossRef]
  54. Yong, G.; Wang, Z.; Chen, J. Two-dimensional and three-dimensional nickel(II) supramolecular complexes based on the new chelating ligand N-(4-carboxyphenyl)iminodiacetic acid: Hydrothermal synthesis and crystal structures. J. Mol. Struct. 2004, 707, 223–229. [Google Scholar] [CrossRef]
  55. CrysAlisPro Software System, 1.171.39.46; Rigaku Corporation: Oxford, UK, 2018.
  56. Sheldrick, G.M. SHELXT—Integrated space-group and crystal-structure determination. Acta Crystallogr. Sect. A Found. Adv. 2015, 71, 3–8. [Google Scholar] [CrossRef] [PubMed]
  57. Sheldrick, G.M. Crystal structure refinement with SHELXL. Acta Crystallogr. Sect. C Struct. Chem. 2015, 71, 3–8. [Google Scholar] [CrossRef] [PubMed]
  58. 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]
  59. Addison, A.W.; Rao, T.N.; Reedijk, J.; van Rijn, J.; Verschoor, G.C. Synthesis, structure, and spectroscopic properties of copper(II) compounds containing nitrogen–sulphur donor ligands; the crystal and molecular structure of aqua [1,7-bis(N-methylbenzimidazol-2′-yl)-2,6-dithiaheptane]copper(II) perchlorate. J. Chem. Soc. Dalton Trans. 1984, 1349–1356. [Google Scholar] [CrossRef]
  60. Choquesillo-Lazarte, D.; Covelo, B.; González-Pérez, J.M.; Castiñeiras, A.; Niclós-Gutiérrez, J. Metal chelates of N-(2-pyridylmethyl)iminodiacetate(2-) ion (pmda). Part I. Two mixed-ligand copper(II) complexes of pmda with N,N-chelating bases. Synthesis, crystal structure and properties of H2pmda·0.5H2O, [Cu(pmda)(pca)]·3H2O (pca=α-picolylamine) and [Cu(pmda)(Hpb)]·5H2O (Hpb=2-(2′-pyridyl)benzimidazole). Polyhedron 2002, 21, 1485–1495. [Google Scholar] [CrossRef]
  61. Román-Alpiste, M.; Martín-Ramos, J.; Castiñeiras-Campos, A.; Bugella-Altamirano, E.; Sicilia-Zafra, A.; González-Pérez, J.; Niclós-Gutiérrez, J. Synthesis, XRD structures and properties of diaqua(iminodiacetato)copper(II), [Cu(IDA)(H2O)2], and aqua(benzimidazole)(iminodiacetato)copper(II), [Cu(IDA)(HBzIm)(H2O)]. Polyhedron 1999, 18, 3341–3351. [Google Scholar] [CrossRef]
  62. Bugella-Altamirano, E.; González-Pérez, J.M.; Choquesillo-Lazarte, D.; Carballo, R.; Castiñeiras, A.; Niclós-Gutiérrez, J. A structural evidence for the preferential coordination of the primary amide group versus the unionised carboxyl group: Synthesis, molecular and crystal structure, and properties of [Cu(HADA)2], a new copper(II) bis-chelate (H2ADA=N-(2-carbamoylmethyl)iminodiacetic acid). Inorg. Chem. Commun. 2003, 6, 71–73. [Google Scholar] [CrossRef]
  63. Macrae, C.F.; Sovago, I.; Cottrell, S.J.; Galek, P.T.A.; McCabe, P.; Pidcock, E.; Platings, M.; Shields, G.P.; Stevens, J.S.; Towler, M.; et al. Mercury 4.0: From visualization to analysis, design and prediction. J. Appl. Crystallogr. 2020, 53, 226–235. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Structure of 2-carboxyphenyliminodiacetic acid.
Figure 1. Structure of 2-carboxyphenyliminodiacetic acid.
Crystals 13 01669 g001
Scheme 1. Synthesis of the complexes 14.
Scheme 1. Synthesis of the complexes 14.
Crystals 13 01669 sch001
Figure 2. (a) Asymmetric unit of the structure of compound 1 with labeling; (b) chain-like structure of the [Cu(cpida)] anion; (c) chain-like structure of the [Cu2(cpida)(H2O)4]+ cation; (d) 1D polymeric structure of compound 1 showing the arrangement of the adjacent chains. Thermal ellipsoids are shown at 50% probability. Hydrogen atoms are omitted for clarity. Color code: O (red), N (blue), carbon (grey), Cu (orange).
Figure 2. (a) Asymmetric unit of the structure of compound 1 with labeling; (b) chain-like structure of the [Cu(cpida)] anion; (c) chain-like structure of the [Cu2(cpida)(H2O)4]+ cation; (d) 1D polymeric structure of compound 1 showing the arrangement of the adjacent chains. Thermal ellipsoids are shown at 50% probability. Hydrogen atoms are omitted for clarity. Color code: O (red), N (blue), carbon (grey), Cu (orange).
Crystals 13 01669 g002
Figure 3. (a) Asymmetric unit of the structure of compound 2 with labeling. Thermal ellipsoids are shown at 50% probability. Hydrogen atoms are omitted for clarity. (b) Packing of compound 2, with view along the c axis; hydrogen atoms and solvent crystallization molecules are omitted for clarity.
Figure 3. (a) Asymmetric unit of the structure of compound 2 with labeling. Thermal ellipsoids are shown at 50% probability. Hydrogen atoms are omitted for clarity. (b) Packing of compound 2, with view along the c axis; hydrogen atoms and solvent crystallization molecules are omitted for clarity.
Crystals 13 01669 g003
Figure 4. (a) Asymmetric unit of compound 3 with labeling. Thermal ellipsoids are shown at 50% probability; (b) Zig-zag motif of [Cu(cpida)] fragments in 3. Hydrogen atoms and sodium ions are omitted for clarity; (c) coordination environment of sodium ions; (d) packing of 3 viewed along the b axis. Hydrogen atoms are omitted for clarity.
Figure 4. (a) Asymmetric unit of compound 3 with labeling. Thermal ellipsoids are shown at 50% probability; (b) Zig-zag motif of [Cu(cpida)] fragments in 3. Hydrogen atoms and sodium ions are omitted for clarity; (c) coordination environment of sodium ions; (d) packing of 3 viewed along the b axis. Hydrogen atoms are omitted for clarity.
Crystals 13 01669 g004
Figure 5. (a) Asymmetric unit of the structure of compound 4 with labeling. Thermal ellipsoids are shown at 50% probability. (b) Zig-zag motif of [Cu(cpida)] fragments in 4. Hydrogen atoms and sodium ions are omitted for clarity. (c) Coordination environment of sodium ions. (d) Packing of compound 4, with view along the c axis. Hydrogen atoms are omitted for clarity.
Figure 5. (a) Asymmetric unit of the structure of compound 4 with labeling. Thermal ellipsoids are shown at 50% probability. (b) Zig-zag motif of [Cu(cpida)] fragments in 4. Hydrogen atoms and sodium ions are omitted for clarity. (c) Coordination environment of sodium ions. (d) Packing of compound 4, with view along the c axis. Hydrogen atoms are omitted for clarity.
Crystals 13 01669 g005
Table 1. Crystallographic data for compounds 14.
Table 1. Crystallographic data for compounds 14.
Compound1234
CCDC number2304883230488123048802304882
Empirical formulaC22H30Cu3N2O19C23H29Cu3N2O17.5C11H8CuNNaO6C11H12CuNNaO8
Formula weight817.10804.10336.71372.75
Temperature (K)293(2)293(2)150(2)150(2)
Wavelength0.71073 Å1.54184 Å1.54184 Å1.54184 Å
Crystal systemMonoclinicMonoclinicOrthorhombicOrthorhombic
Space groupP 21/nP 21/nPbcaIba2
a (Å)18.0990 (7)18.1311 (4)14.4270 (8)28.1538 (16)
b (Å)9.7523 (3)9.7972 (2)6.4340 (4)14.5726 (7)
c (Å)18.2152 (6)18.1195 (5)23.3780 (16)6.2930 (3)
α (°)90909090
β (°)93.438 (3)94.229(2)9090
γ (°)90909090
Volume (Å3)3209.32 (19)3209.88 (13)2170.0 (2)2581.9 (2)
Z4488
Dcalc(g·cm−3)1.6911.6642.0611.918
Absorption coefficient (mm−1)2.0533.0113.5403.164
F(000)1660163213521512
Crystal dimensions (mm)0.22 × 0.17 × 0.110.43 × 0.30 × 0.220.19 × 0.07 × 0.030.33 × 0.11 × 0.05
Theta range for data collection (°)1.636 to 26.3725.135 to 66.5854.869 to 67.0335.607 to 66.556
Index ranges−14 ≤ h ≤ 22−21 ≤ h ≤ 19−17 ≤ h ≤ 16−33 ≤ h ≤ 27,
−12 ≤ k ≤ 12−11 ≤ k ≤ 8−7 ≤ k ≤ 4−12 ≤ k ≤ 17
−21 ≤ l ≤ 22−18 ≤ l ≤ 21−27 ≤ l ≤ 12−7 ≤ l ≤ 4
Reflections collected13,13512,30846662980
Independent reflections, R(int)6552, 0.03675452, 0.01631927, 0.02841599, 0.0206
Data/restraints/parameters6552/3/4375452/0/4211927/0/1811599/1/181
Goodness of fit on F21.0221.0911.1731.081
Final R indices [I > 2sigma(I)]a,bR1 = 0.0485, wR2 = 0.1251R1 = 0.0431, wR2 = 0.1338R1 = 0.0571, wR2 = 0.1406R1 = 0.0255, wR2 = 0.0679
R indices (all data)a,bR1 = 0.0649, wR2 = 0.1375R1 = 0.0451, wR2 = 0.1353R1 = 0.0643, wR2 = 0.1438R1 = 0.0259, wR2 = 0.0682
Largest diff. peak and hole (e.Å−3)0.90/−0.541.65/−0.880.81/−0.820.37/−0.41
a R1 = Σ||F0|−|Fc||/Σ|Fc|, b wR2 = (Σ[w(F02−Fc2)2]/Σ[w(F02)2)1/2.
Table 2. Selected bond distances (Å) for the coordination center in compounds 1 and 2 .
Table 2. Selected bond distances (Å) for the coordination center in compounds 1 and 2 .
Compound 12
Distances
Cu(1)–O(1)1.918(3)Cu(1)–O(1)1.917(2)
Cu(1)–O(6)#11.936(3)Cu(1)–O(6)#31.933(2)
Cu(1)–O(2)#12.239(3)Cu(1)–O(2)#32.247(2)
Cu(1)–O(3)#11.943(3)Cu(1)–O(3)#31.935(3)
Cu(1)–N(1)#12.043(3)Cu(1)–N(1)#32.043(3)
Cu(2)–O(7)1.911(3)Cu(2)–O(8)#42.238(2)
Cu(2)–O(8)#22.251(3)Cu(2)–O(7)1.905(2)
Cu(2)–O(9)#21.975(3)Cu(2)–O(12)#41.976(3)
Cu(2)–O(12)#21.959(3)Cu(2)–O(9)#41.983(3)
Cu(2)–N(2)#22.028(3)Cu(2)–N(2)#42.036(3)
Cu(3)–O(2W)1.975(3)Cu(3)–O(2W)1.970(3)
Cu(3)–O(5)1.976(3)Cu(3)–O(5)1.990(3)
Cu(3)–O(1W)2.179(3)Cu(3)–O(4W)1.994(3)
Cu(3)–O(4W)1.984(3)Cu(3)–O(3W)2.179(3)
Cu(3)–O(3W)1.951(3)Cu(3)–O(1W)1.946(3)
Symmetry codes: #1 −x + 1/2, y + 1/2, −z + 3/2; #2 −x + 1/2, y − 1/2, −z + 5/2; #3 −x + 3/2, y + 1/2, −z + 3/2; #4 −x + 1/2, y − 1/2, −z + 3/2.
Table 3. Selected bond distances (Å) for the coordination center in compounds 3 and 4 .
Table 3. Selected bond distances (Å) for the coordination center in compounds 3 and 4 .
Compound 34
Distances
Cu(1)–O(1)#51.935(4)Cu(1)–O(5)1.903(2)
Cu(1)–O(5)1.892(4)Cu(1)–O(3)1.947(2)
Cu(1)–O(3)1.940(4)Cu(1)–O(1)#91.952(2)
Cu(1)–N(1)2.046(5)Cu(1)–O(1)2.390(3)
Na(1)–O(1)2.706(5)Cu(1)–N(1)2.054(3)
Na(1)–O(5)2.365(4)Na(1)–O(5)2.327(3)
Na(1)–O(3)#62.903(4)Na(1)–O(3)#102.919(3)
Na(1)–O(3)#72.408(4)Na(1)–O(3)#112.371(3)
Na(1)–O(6)#82.361(5)Na(1)–O(2)#92.809(3)
Na(1)–O(4)#62.397(5)Na(1)–O(6)#122.362(3)
Na(1)–O(2)#52.690(5)Na(1)–O(4)#102.401(3)
Na(1)–O(1)2.701(3)
Symmetry codes: #5 −x + 1, y + 1/2, −z + 1/2; #6 x + 1/2, y, −z + 1/2; #7 −x + 1, y − 1/2, −z + 1/2; #8 −x + 3/2, y − 1/2, z; #9 −x + 1/2,−y + 1/2, z − 1/2; #10 −x + 1/2, y + 1/2, z; #11 −x + 1/2, −y + 1/2,z + 1/2; #12 x, −y + 1, z + 1/2.
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

Martínez, S.; Kremer, C.; González-Platas, J.; Mendoza, C. New Polynuclear Coordination Compounds Based on 2–(Carboxyphenyl)iminodiacetate Anion: Synthesis and X-rays Crystal Structures. Crystals 2023, 13, 1669. https://doi.org/10.3390/cryst13121669

AMA Style

Martínez S, Kremer C, González-Platas J, Mendoza C. New Polynuclear Coordination Compounds Based on 2–(Carboxyphenyl)iminodiacetate Anion: Synthesis and X-rays Crystal Structures. Crystals. 2023; 13(12):1669. https://doi.org/10.3390/cryst13121669

Chicago/Turabian Style

Martínez, Sebastián, Carlos Kremer, Javier González-Platas, and Carolina Mendoza. 2023. "New Polynuclear Coordination Compounds Based on 2–(Carboxyphenyl)iminodiacetate Anion: Synthesis and X-rays Crystal Structures" Crystals 13, no. 12: 1669. https://doi.org/10.3390/cryst13121669

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