Next Article in Journal
8-oxoguanine and 8-oxodeoxyguanosine Biomarkers of Oxidative DNA Damage: A Review on HPLC–ECD Determination
Previous Article in Journal
Supported Ionic Liquids Used as Chromatographic Matrices in Bioseparation—An Overview
Previous Article in Special Issue
Structural Identification of Metalloproteomes in Marine Diatoms, an Efficient Algae Model in Toxic Metals Bioremediation
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Confirming the Molecular Basis of the Solvent Extraction of Cadmium(II) Using 2-Pyridyl Oximes through a Synthetic Inorganic Chemistry Approach and a Proposal for More Efficient Extractants

by
Anastasia Routzomani
1,
Zoi G. Lada
1,2,
Varvara Angelidou
1,
Catherine P. Raptopoulou
3,
Vassilis Psycharis
3,*,
Konstantis F. Konidaris
4,*,
Christos T. Chasapis
2,5,* and
Spyros P. Perlepes
1,2,*
1
Department of Chemistry, University of Patras, 265 04 Patras, Greece
2
Institute of Chemical Engineering Sciences (ICE-HT), Foundation for Research and Technology-Hellas (FORTH), Platani, P.O. Box 1414, 265 04 Patras, Greece
3
Institute of Nanoscience and Nanotechnology, NCSR “Demokritos”, Aghia Paraskevi, Attikis, 153 10 Athens, Greece
4
Department of Science and High Technology and INSTM, University of Insubria, 22 100 Como, Italy
5
NMR Facility, Instrumental Analysis Laboratory, School of Natural Sciences, University of Patras, 265 04 Patras, Greece
*
Authors to whom correspondence should be addressed.
Molecules 2022, 27(5), 1619; https://doi.org/10.3390/molecules27051619
Submission received: 4 February 2022 / Revised: 22 February 2022 / Accepted: 24 February 2022 / Published: 28 February 2022
(This article belongs to the Special Issue Metal Intoxication: General Aspects and Chelating Agents)

Abstract

:
The present work describes the reactions of CdI2 with 2-pyridyl aldoxime (2paoH), 3-pyridyl aldoxime (3paoH), 4-pyridyl aldoxime (4paoH), 2-6-diacetylpyridine dioxime (dapdoH2) and 2,6-pyridyl diamidoxime (LH4). The primary goal was to contribute to understanding the molecular basis of the very good liquid extraction ability of 2-pyridyl ketoximes with long aliphatic chains towards toxic Cd(II) and the inability of their 4-pyridyl isomers for this extraction. Our systematic investigation provided access to coordination complexes [CdI2(2paoH)2] (1), {[CdI2(3paoH)2]}n (2), {[CdI2(4paoH)2]}n (3) and [CdI2(dapdoH2)] (4). The reaction of CdI2 and LH4 in EtOH resulted in a Cd(II)-involving reaction of the bis(amidoxime) and isolation of [CdI2(L’H2)] (5), where L’H2 is the new ligand 2,6-bis(ethoxy)pyridine diimine. A mechanism of this transformation has been proposed. The structures of 1, 2, 3, 2EtOH and 5 were determined by single-crystal X-ray crystallography. The complexes have been characterized by FT-IR and FT-Raman spectra in the solid state and the data are discussed in terms of structural features. The stability of the complexes in DMSO was investigated by 1H NMR spectroscopy. Our studies confirm that the excellent extraction ability of 2-pyridyl ketoximes is due to the chelating nature of the extractants leading to thermodynamically stable Cd(II) complexes. The monodentate coordination of 4-pyridyl ketoximes (as confirmed in our model complexes with 4paoH and 3paoH) seems to be responsible for their poor performance as extractants.

Graphical Abstract

1. Introduction

Organic matter and heavy toxic metals are the main pollutants of wastewaters, the threat from the latter being more serious [1,2,3,4,5,6]. This is due to the non-biodegradable and non-decomposable nature of the toxic metals, making the development of efficient approaches for their removal and uptake extremely important [7]. Methods in action involve chemical precipitation, microbial treatment, electrodeposition, reverse osmosis, physical/chemical adsorption and solvent extraction [1,8,9,10,11,12]. Solvent extraction is very useful for base metal recovery; the desired metal ions are transferred selectively (after the ore is leached into an aqueous solution) to an H2O-immiscible phase with ligands as ion-exchangers, which release an equivalent number of H3O+ ions back to the aqueous feed solution [13]. For the removal of toxic metal ions from aqueous environments, a special method of solvent extraction, named liquid-liquid extraction, is common [12]. Three types of liquid-liquid extraction are currently used. All three involve the metal-ion association with an organic complexant (extractant) to form a species that is transferred from the aqueous to the organic phase in a two-phase system [12]. In the first type, both the complexant and the metal ion are soluble in the organic phase. In the second type, the complexant and the metal ion source are insoluble in the aqueous phase, the complexation reaction occurs at the interphase region and the metal species is then transferred into the organic phase. In the third type of liquid-liquid extraction, the extractant is soluble in the organic phase (hexane, kerosene, and chloroform for laboratory experiments) and the source of the metal ion is soluble only in the aqueous phase; after the complexation reaction that occurs at the interphase surface, the metal-extractant coordination complex is transferred into the organic phase. The present model study refers to the latter type of liquid-liquid extraction. Most efficient extractants include chelating or even macrocyclic ligands [12]. Today, it is firmly established that an effective extractant should [14]: (a) selectively coordinate to the toxic metal ion having no or very weak affinity for alkali and alkaline metal ions (e.g., Na+, Ca2+, etc.) which are present at higher concentrations in waste and natural waters; (b) give somewhat thermodynamically stable complex with the toxic metal ion; (c) have fast kinetics with the metal ion to be extracted; (d) resist hydrolysis; and (e) result in a reversible complexation allowing for the easy and complete recovery of the metal ion without destruction or decomposition of the extractant.
The history of cadmium was brilliant in the past [15,16]. Contrary to the past, the present and future of cadmium (both as element and +2 cation) seem dark. Cd is a very toxic metal and is considered as one of the 13 most dangerous metals by the Environmental Protection Agency (EPA) in the U.S. [17,18]. Its ion has generally a short life span and is rendered inert after a period of action. The maximum limit of Cd(II) for humans in H2O is 10 mg L−1. It is introduced into the environment by metallurgical processes (e.g., Pb-Zn mining) and wastes from electroplating and companies producing pigments, materials for photography and alkaline batteries [1,2,15,19,20,21]. Exposure to Cd(II) causes damages in heart, bones, lungs, and mainly in the kidneys where it is collected affecting their filtering ability [22]. Cadmium(II) compounds are widely classified as carcinogens for humans; most data come from detailed studies which have proven that occupational exposure to Cd(II) sources is closely associated with lung cancer, and possibly to human prostate and renal cancers [23]. Thus, methods for Cd(II) uptake from wastewaters are becoming more and more important.
In the liquid-liquid extraction of toxic Cd(II), several extractants have been used including EDTA derivatives, organophosphorus compounds, dithiocarbamates, crown ethers, calix[4]arene derivatives, 8-quinolols, pyridine carboxamides and pyridyl oximes [12,24]. The stimulus of the present work and a previous publication from our groups [14] was an excellent liquid-liquid extraction study of Cd(II) from aqueous chloride solutions using 1-(2-pyridyl)-tridecane-1-one (2PC12), 1-(2-pyridyl)-pentadecane-1-one (2PC14), 1-(4-pyridyl)-tridecane-1-one (4PC12) and 1-(4-pyridyl)-pentadecane-1-one (4PC14) oximes [25], Figure 1. We are doing a parenthesis here to mention that 2PC12, among other 2-pyridyl ketoximes, has also been successfully used for the liquid-liquid extraction of copper(II) from aqueous solutions containing Cl ions [26]. Parus, A. et al. examined carefully the influence of extractant, Cd(II) and Cl concentrations, and the nature of several polar and non-polar organic solvents (diluents) on the extraction efficiency. The study revealed two general experimental facts: (a) Cd(II) is extracted using only 2PC12 and 2PC14. 4PC12 and 4PC14 did not transport complexes with Cd(II) to organic phases. The organic phase employed was CHCl3 or hydrocarbons mixed with decan-1-ol. The metal ion was stripped from the loaded organic phase with H2O and aqueous NH3 solutions; and (b) Based on the effect of the varying concentrations of 2PC12 and 2PC14 on the extraction capability, it was proposed that the chemical species formed and transferred into the organic phase are the neutral complexes [CdCl2(2PC12)2] and [CdCl2(2PC14)2]. Using a synthetic coordination chemistry approach [14], we proved that such complexes are capable of existence. The primary goal of the present study was to contribute in understanding the molecular basis of the experimental fact (a) mentioned above. Although the reason of the superior extraction capability of 2PC12, 2PC14 compared to those of 4PC12, 4PC14 might be obvious, i.e., the formation of stable 5-membered chelating rings with the N-donor atoms of the former extractants—which give a thermodynamic stability of the complexes in solution—and the inability of chelating behavior in the case of 4PC12, 4PC14, we were interested in providing synthetic and structural evidence for this working with model complexes. Thus, we used the three isomers of pyridyl aldoximes (2-pyridyl aldoxime, 2paoH; 3-pyridyl aldoxime, 3paoH; 4-pyridyl aldoxime, 4paoH), Figure 1, which gave crystalline complexes. The compounds 2paoH and 4paoH are somewhat satisfactory analogs (albeit not ideal ones) of the extractants 2PC12, 2PC14 and 4PC12, 4PC14, respectively. The main difference is the presence of a long aliphatic (hence hydrophobic) chain in the real extractants instead of the -CHNOH group in the ligands. Cadmium(II) complexes with the ligand 3paoH were also studied. This paper describes results from the synthetic investigation of the general reaction systems CdI2/2paoH, CdI2/3paoH and CdI2/4paoH. For consistency reasons, we used the same inorganic anion of the Cd(II) source; although we originally worked with CdCl2·2H2O to better simulate the real extraction conditions (e.g., chloride solutions), CdI2 gave better crystallinity of the products and their structures were solved through single crystal X-ray crystallography (vide infra).
A secondary goal of this work was to propose types of ligands that might be better extractants than the 2-pyridyl ketoximes 2PC12 and 2PC14. For this reason, we performed reactions of CdI2 with 2,6-diacetylpyridine dioxime (dapdoH2, Figure 1) and 2,6-diacetylpyridine diamidoxime (LH4, Figure 1), to investigate whether the potentially N(pyridyl), {N(oxime)}2 tridentate chelating character of these ligands could be realized; in such a case, molecules analogous to dapdoH2 and LH4 might be better extractants for Cd(II). The ligand LH4 possesses two amidoxime moieties which, when introduced in materials (e.g., polymers) give excellent adsorbents for the efficient recovery of Cd(II) from aqueous media [1,21,28].
The spectroscopic studies and the structures of the prepared complexes emphasize the importance of the information emerging from synthetic investigations that otherwise would not have been revealed through exclusively solution studies, while simultaneously shedding light onto the connection of the requisite chemistry with the removal of toxic Cd(II) using pyridyl oxime extractants [19]. The present work can be considered as a continuation of our intense interest in the coordination chemistry of pyridyl aldo/ketoximes and dioximes, and amidoximes/amidodioximes [14,27,29,30,31,32,33,34,35,36,37,38,39,40] and on various aspects of Cd(II) chemistry [14,40,41,42,43,44,45].

2. Results and Discussion

2.1. Comments on the Syntheses of the Complexes

Since the pH of the aqueous phases during the real extraction experiments was ~4.0 and the extractants were neutral, we avoided adding external bases (e.g., OH, Et3N, etc.) which would deprotonate the ligands used. The optimized reaction systems used are conveniently summarized in Figure 2.
The reaction schemes illustrated in Figure 2 deserve synthetic comments. Thus: (i) The use of 1:1 reaction ratio in the case of the CdI2/(2,3,4)paoH systems (the complexes were isolated using 1:2 reaction ratios, which coincide with their stoichiometries) does not affect the product identity, leading again to compounds 13 (microanalytical and IR evidences). (ii) Complexes 2,3 can be precipitated from MeCN solutions (IR evidence), but their crystallinity is lower compared to that using EtOH and Me2CO as solvents, respectively. (iii) The use of excess dapdoH2 in its reaction with CdI2 (e.g., a 2:1 dapdoH2:CdI2 ratio) does not affect the product identity (IR and Raman evidences); the complex was originally isolated using the 1:1 reaction which coincides with its stoichiometry. However, if the reaction solution is concentrated enough (to increase the low yield), the solid product is contaminated with free ligand; and (iv) The study of the 1:1 CdI2/LH4 reaction system in EtOH provided us with surprising results. The initially precipitated solid (in a yield of ~10%) was analyzed perfectly as [CdI2(LH4)] (5a), as expected. The identity of the product was confirmed by its IR and 1H NMR spectra (vide infra). Despite our intense efforts, we could not grow crystals of 5a for detailed structural characterization. Somewhat to our surprise, an unexpected experimental result was observed during the storage of the filtrate at room temperature; within four days, its color turned slowly pale violet! The pale violet solution was kept in the refrigerator (~5 °C) and X-ray quality, colorless crystals of [CdI2(L′H2)] (5) were grown in a ~40% yield (based on the initially used metal source) within one week. Single-crystal X-ray crystallography revealed that 5 contained the transformed ligand L′H2, i.e., the LH4 → L′H2 transformation had occurred. The free compound L′H2 is not known in organic chemistry. We do believe that the observed transformation is CdI2-promoted/assisted, since ethanolic solutions of LH4 are stable for months. The metal ion-involving reactions of the amidoxime group are well known [46] in the field of the reactivity of coordinated ligands. No analogous reaction has been observed to date in the chemistry of amidoximes. Without any detailed mechanistic studies, we propose the simplified mechanism shown in Figure 3 for the LH4 → L′H2 transformation. The formation of the L′H2 probably involves nucleophilic attack of the solvent to the oxime C atom (Cd(II) coordination to the oxime nitrogen might activate the amidoxime groups towards nucleophilic attack) followed by the reduction in the oxime group by the generated HI, with the simultaneous production of I2; the latter justifies the observation of the pale violet color of the reaction solution. In accordance with our proposal, the transformation does not take place in MeCN from which a very small quantity of 5a is precipitated.

2.2. Spectroscopic Characterization in Brief

The complexes were characterized in the solid state by IR and Raman spectroscopies. Representative spectra are shown in Figure 4, Figure 5, Figure S1, Figure S2, Figure S3, Figure S4 and Figure S5. The spectra do not exhibit bands that are present in the free ligands 2paoH, 3paoH, 4paoH, dapdoH2 and LH4, suggesting their purity; if such bands were present, the complexes would have been contaminated with the free ligands used as starting materials. The presence of neutral oxime groups in 1, 2, 3 and 4 (well-dried unsolvated sample) is manifested by broad bands (with 2–4 submaxima in the spectra of 1 and 4) at ~3400 cm−1 assigned to ν(OH) [38,39]. The bands at 3444 and 3292 cm−1 in the IR spectrum of 5 reflect the existence of the imino (=NH) groups in the complex [47]. In 5a, the bands at 3480 [νas(NH2)], 3432 [νs(NH2)] and 3372 [ν(OH)] reflect the existence of -NH2 and -OH groups supporting the incorporation of coordinated LH4 in the complex. The corresponding bands in the free LH4 appear at 3484, 3420 and ~3380 cm−1. The broadness of the ν(OH), ν(NH), νas(NH2) and νs(NH2) bands, combined with their relatively low wavenumber, are both indicative of hydrogen bonding [38,39]. As expected, the O-H and N-H stretching vibrations are hardly seen in the Raman spectra of the complexes. The peaks at 3076-2922 cm−1 are assigned to ν(C-H) vibrations [14,48,49]. The in-plane, δ(py), and out-of-plane, γ(py), deformation vibrations of the 2-pyridyl ring of free paoH (at 627 and 404 cm−1, respectively) shift upwards (at 646 and 466 cm−1, respectively) in 1 suggesting coordination of the ring-N atom [38]. The δ(py) vibration appears as a weak peak in its Raman spectrum [48]. The same trend is observed in the vibrational spectra of the other complexes. For example, the δ(py) and γ(py) bands are at 658 and 486 cm−1, respectively, in the IR spectrum of 3, while the corresponding vibrations in the spectrum of free 4paoH are located at 640 and 452 cm−1, respectively. The ν(C=N) vibration of the oxime group(s) in the IR spectra of 1, 2, 3, 4 and 5a appear as medium to weak bands at 1640, 1624, 1638, 1594 (overlap** with an aromatic stretch) and 1654 cm−1, respectively [38]; the corresponding Raman vibrations are assigned [14,48] to the peaks at 1632 (1), 1624 (2), 1622 (3) and 1589 (4) cm−1. The wavenumbers for 2 and 3 are approximately the same with those of the free ligands 3paoH and 4paoH, respectively. This is strong evidence that the oxime nitrogen does not participate in coordination to Cd(II), a fact that has been confirmed through single-crystal X-ray crystallography (vide infra). The ν(C=N) band/peak (IR at 1594 cm−1 and Raman at 1638 cm−1) for 4 is located at lower wavenumbers compared with those of free dapdoH2, suggesting oxime-N coordination. Somewhat to our surprise, the 1626 cm−1 band of 2paoH is shifted to a higher wavenumber in the IR spectrum of 1 (1640 cm−1) for which the oxime-N coordination has been confirmed (vide infra). This fact, which is not unusual [14], has been interpreted [50] on the basis that some ligands containing a C=N bond (with the carbon atom attached to an aromatic ring) have been shown to undergo a change in the s character of the N lone pair upon coordination; the s character of the N orbital in the C=N bond increases resulting in a greater C=N stretching force constant relative to the free neutral ligands, and this in turn shifts the ν(C=N) band in the spectra of the complexes to higher frequencies. An analogous trend is observed in the IR spectrum of 5a for which coordination of both oxime N atoms is proposed. The medium-intensity IR band at 942 cm−1 for free 2paoH is assigned to the ν(NO)oxime mode; the corresponding weak Raman peak appears at 950 cm−1 [49]. The 942 cm−1 band is shifted to a lower wavenumber (932 cm−1) in 1; this shift is due [14,39] to the coordination of the neutral oxime nitrogen. The same trend is observed in the Raman spectrum of the complex where this mode is located at 925 cm−1. The assignment of the ν(NO)oxime mode for 3paoH, 4paoH, dapdoH2 and LH4 is not an easy task and any discussion about coordination (or non-coordination) shifts in the spectra of the complexes would be risky.
The complexes were characterized in solution by 1H NMR spectroscopy in d6-DMSO (Figure S6, Figure S7 and Figure S8). The spectrum of free 2paoH displays singlet signals at δ 11.66 and 8.08 ppm assigned to the hydroxyl proton and to the proton attached to the oxime carbon, respectively, and a doublet signal at δ 8.57 ppm assigned to the α aromatic proton (i.e., the proton bonded to the carbon next to the pyridyl nitrogen atom). The corresponding signals in the spectrum of 1 appear at δ 10.97, 7.42 and 9.13 ppm. The upfield shift of the oxime carbon-bonded proton and the downfield shift of the α proton in the spectrum of the complex both indicate coordination of the two 2paoH N atoms in solution [51], suggesting that the structure of 1 is retained in solution. This is corroborated by the almost zero value of the molar conductivity of the complex in DMSO (ΛM = 4 S cm2 mol−1 for a 10−3 M solution at 25 °C) [52]. Given the stability of 1 in DMSO, we were rather surprised to realize that 4 (which possesses two chelating rings per ligand dapdoH2) decomposes in solution! Its 1H NMR spectrum in d6-DMSO is identical to that of free dapdoH2 displaying a simple set of signals at δ 11.51 (-OH, singlet), 7.80 (aromatic protons, multiplet) and 2.25 (-CH3, singlet) ppm in the expected 2:3:6 integration ratio. This result, together with the negligible ΛM value in DMSO, indicates a decomposition probably through Equation (1), where x ≥ 4. The CdII-I bond is very stable as suggested by the absence of crystal structures of Cd(II)-DMSO complexes possessing ionic iodides [53]. A rather poor-quality (due to solubility reasons) 1H NMR spectrum of 4 in CDCl3 is complicated indicating two different solution species, both of which seem to contain coordinated dapdoH2. The spectrum of 2 in d6-DMSO is identical to that of free 3paoH displaying signals at δ 11.57 (-OH, singlet), 8.21 (-CHNOH) and doublets/multiplets at 8.75, 8.57, 8.00, 7.45 (aromatic protons) ppm in the expected 1:1:4 integration ratio. Similarly, the spectrum of 3 in the same solvent shows signals at δ 11.82 (-OH, singlet), 8.18 (-CHNOH, singlet), and 8.60, 7.55 (aromatic protons) ppm being identical with that of free 4paoH. These data indicate that complexes 2 and 3 decompose in solution releasing 3paoH and 4paoH, respectively. The ΛM values of 2 and 3 in DMSO were very small probably indicating decomposition through Equation (2), x ≥ 4 and a = 2 or 3.
[ C d I 2 ( d a p d o H 2 ) ] + x D M S O DMSO [ C d I 2 ( D M S O ) x ] + d a p d o H 2
{ [ C d I 2 ( a p a o H ) 2 ] } n + xn   DMSO DMSO n   [ C d I 2 ( D M S O ) 2 ] + 2 n   apaoH  

2.3. Description of Structures

The structures of 13, 2EtOH and 5 were determined by single-crystal X-ray crystallography. Crystallographic data are gathered in Table 1. Structural plots are shown in Figure 6, Figure 7, Figure 8, Figure 9, Figure 10 and Figure 11 and S9–S12. Selected interatomic distances and angles are listed in Table 2, Table 3, Table 4, Table 5 and Table 6.
Complex [CdI2(2paoH)2] (1) crystallizes in the monoclinic space group C2/c. As the complex possesses a 2-fold axis of symmetry passing through Cd1 and bisecting the I1-Cd1-I1 angle (Wyckoff position 4e: O, y, 1/4), there is 1/2 [CdI2(2paoH)2] molecule in the asymmetric unit of the cell. The CdII center forms coordination bonds with two terminal iodo (or iodido) groups (I1, I1 ′), two oxime nitrogen atoms (N1, N1 ′) and two 2-pyridyl nitrogen atoms (N2, N2 ′); the four nitrogen atoms belong to two 1.011 (Harris notation [54]) 2paoH ligands. The coordination polyhedron of the metal ion is a very distorted octahedron, the cis and trans donor atom-metal ion-donor atom bond angles being in the ranges 66.9(2)–114.8(1) and 145.8(2)–153.6(1) °, respectively. The distorted geometry arises from the small bite angles of the two 5-membered N(oxime)CCN(2-pyridyl)Cd1 rings; these two angles (equal by symmetry) are 66.9(2) °, much smaller than the ideal value of 90 °. The octahedral molecule is the cis-cis-trans isomer considering the octahedral positions of the iodo groups, the 2-pyridyl and the oxime nitrogen atoms, respectively. The CdII-N(2-pyridyl) bond length [2.402(5) Å] is smaller than the CdII-N(oxime) one [2.457(6) Å]. There are two intramolecular hydrogen bonds with the uncoordinated oxime oxygen atoms (O1, O1 ) as donors and the coordinated iodo groups (I1, I1 ) as acceptors, namely O1-H(O1)I1 [and O1 -H(O1 )I1]. Their parameters are: O1I1 = 3.605(5) Å, H(O1)...I1 = 2.90 Å and O1-H(O1)I1 = 142.0 °.
Neighboring molecules of 1 interact through π-π stacking interactions involving the 2-pyridyl rings (symmetry operation: −x −1/2, −y + 1/2, −z) forming chains parallel to the [101] crystallographic direction (Figure S9), which are further connected through weak hydrogen bonding interactions creating the 3D architecture of the crystal structure. The distance between the neighboring centrosymmetric 2-pyridyl rings within the chain is 3.83(1) Å.
Compound 2 consists of linear 1D chains. It crystallizes in the monoclinic space group C2/m and the asymmetric unit contains ¼ of the repeating unit [CdI2(3paoH)2]. The Cd1 centers sit on Wyckoff positions 2a (0, 0.0, 0) with 2/m point group symmetry, with the 2-fold axis being parallel to the b axis and the mirror plane cutting it at the y = 0.0 point. The atoms of the organic ligand are located on a mirror plane as all occupy the 4i (x, 0.0, z) Wyckoff positions with point group symmetry m. The I1 atoms are also located on 4i-type sites (x, 0, y) with a point group symmetry m, but in this case the mirror plane crosses the b axis at the y = 0.5 point. The CdII and I atoms form chains parallel to the b axis and the 3paoH ligands are bonded to the metal ions in directions normal to the chain axis. Thus, the CdII atoms are doubly bridged by two symmetric μ-iodo groups and two monodentate 1.010 3paoH ligands complete a slightly distorted octahedral coordination at each metal. The donor atom of 3paoH is the pyridyl nitrogen. The Cd-I bond lengths in 2 [2.986(1) Å] are larger than in 1 [2.829(1) Å] due to the bridging character of the iodo ligands in the former as compared to their terminal character in the latter. The CdII…CdII distance is 4.162(1) Å. Neighboring chains in the crystal interact through intermolecular O1-H(O1)I1 hydrogen bonds and form layers parallel to the (110) plane (Figure S10). The metric parameters of the crystallographically unique hydrogen bond O1-H(O1)I1 (x + 1/2, y + 1/2, z) are: O1I1 = 3.528(10) Å, H(O1)I1 = 2.63(3) Å and O-H(O1)I1 = 174(12)°.
The molecular structure and the supramolecular features of 3 are strikingly similar with those of 2. The Cd1 centers sit on the Wyckoff positions 2b (0, 0.5, 0) with 2/m point group symmetry; the 2-fold axis is parallel to the b axis and the mirror plane cuts it at the y = 0.5 point. Again, the atoms of the 1.010 4paoH ligand are located on a mirror plane, as all occupy the 4i (x, 0.5, z) Wyckoff positions with m point group symmetry. The I1 atoms are also located on 4i-type sites (x, 0, y) with point group symmetry m, but in this case the mirror plane crosses the b axis at the y= 0.0 point. The CdII…CdII distances are exactly the same [4.162(1) Å] in 2 and 3. Similarly to the crystal structure of 2, neighboring chains of 3 interact through O(oxime)-HI hydrogen bonds forming layers parallel to the (110) plane (Figure S11). The metric parameters are somewhat different compared to those in 2, due to the different positions of the oxime groups in the pyridyl rings; these are O1-H(O1)I1 (x − 1/2, y + 1/2, z) = 3.501(3) Å, H(O1)I1 (x − 1/2, y + 1/2, z) = 2.81(5) Å and O1-H(O1)I1 (x − 1/2, y + 1/2, z) = 150(4) Å.
The crystal structure of 4·2EtOH consists of mononuclear molecules [CdI2(dapdoH2)] and lattice EtOH molecules in an 1:2 ratio. The asymmetric unit of the cell contains the full complex molecule and the two solvent molecules. Each of the latter interacts (as acceptor) with one “free” (i.e., uncoordinated) oxime oxygen atom (donor) of the dapdoH2 ligand (Figure 9, Table S1). The CdII atom forms coordination bonds with two terminal iodo ligands (I1, I2), the two oxime nitrogen atoms (N1, N3) and the pyridyl nitrogen atom (N2) of dapdoH2. Thus, the organic molecule behaves as a 1.00111 ligand and participates in two 5-membered chelating rings with the metal ion. The CdII-N(pyridyl) bond [2.333(2) Å] is slightly stronger than the CdII-N(oxime) bonds [2.421(2), 2.443(2) Å]. The terminal CdII-I bonds [2.722(1), 2.733(1) Å] are stronger than the corresponding bonds in 1 [2.829(1) Å], and this is due to the lower coordination number of CdII in 4·2EtOH (five) compared with that in 1 (six). The coordination geometry of CdII in the complex is extremely distorted, a fact that is primarily attributed to the small bite angles of the two 5-membered N(oxime)CCN(pyridyl)Cd chelating rings; both N(oxime)-Cd-N(pyridyl) coordination angles are 67.5(1)°. The geometry can be either described as a very distorted trigonal bipyramidal one with atoms N1 and N3 defining the axial positions, or as a very distorted square pyramidal one with atoms I1, I2, N1, N3 occupying the basal plane and N2 being at the apical position.
The molecules [CdI2(dapdoH2)] form pairs through π-π interactions indicated with dashed dark green lines in Figure 10 (the distance between the planes of neighboring, centrosymmetrically-related molecules is 3.81(1) Å, symmetry: −x, −y + 2, −z + 2) and C4-H(C4)-I2 hydrogen bonds. Neighboring pairs interact further through C9-HC(C9)O2 hydrogen bonds along the [110] direction and through π-π interactions along the b axis, thus forming layers parallel to the (001) plane. The distance between the planes of neighboring, centrosymmetrically-related (−x, −y + 1, −z + 2) molecules along the b axis direction is 3.49(1) Å, and this interaction is indicated with dashed light green lines in Figure 10. Metric parameters of the hydrogen bonds are listed in Table S1.
Compound 5 crystallizes in the orthorhombic space group Pcnb. As the complex possesses a 2-fold axis of symmetry passing through Cd1, N2 and C6 atoms (Wyckoff position 4c: 0, ¼, z), the asymmetric unit of the cell contains ½ of the [CdI2(L′H2)] molecule. The 5-coordinate CdII atom forms coordination bonds with the two terminal iodo groups (I1, I1′), the two imino nitrogen atoms (N1, N1′) and the pyridyl nitrogen atom (N2) of the transformed ligand L’H2. Thus, the organic molecule behaves as a 1.00111 ligand participating in two 5-membered chelating rings with the metal center. The terminal CdII-I bonds [2.738(1) Å] are almost identical to those of 4·2EtOH [average 2.727(1) Å], a consequence of the 5-coordination of CdII in the two complexes. As in 4·2EtOH, the coordination polyhedron of the metal ion in 5 is extremely distorted, a fact primarily arising from the small N1-Cd1-N2 and N1 ′-Cd1-N2 [68.2(1)°] coordination angles of the chelating “parts” of L′H2. The polyhedron can be better described as a very distorted trigonal bipyramid with atoms N1 and N1 ′ occupying the axial positions. Neighboring molecules of 5 interact through pairs of C(methyl)-Hπ interactions forming chains parallel to the a axis; neighboring chains interact through C(methyl)-HI hydrogen bonds creating layers parallel to the (001) crystallographic plane (Figure S12, Table S2). Compound 5 is the first structurally characterized complex of any metal containing the new ligand L′H2.
Complexes 2 and 4·2EtOH are the first structurally characterized Cd(II) complexes with 3paoH and dapdoH2 ligands, respectively. Compounds 1 and 3 join a small family of structurally characterized Cd(II) complexes of 2paoH and 4paoH [55,56,57,58,59,60], mainly reported by Fonari’s group in a series of excellent crystal engineering studies. The 2paoH complexes reported are [Cd(O2CMe)2(2paoH)2] [50], [Cd2(suc)(2paoH)4(H2O)2]](BF4)2 [56], {[Cd(suc)(2paoH)2]}n [56], [Cd(HCO2)2(2paoH)2] [59], {[Cd(1,4-bdc)(2paoH)]·1.5DMF}n [59], {[Cd(SO4)(2paoH)(H2O)]}n [59], {[Cd(fum)(2paoH)2]}n [60], {[Cd(1,3-bdc)(2paoH)]}n [60], where suc2− is the succinate(-2) ligand, 1,4-bdc2- is the 1,4-benzenedicarboxylate(-2) ligand, fum2− is the fumarate(-2) ligand and 1,3-bdc2- is the 1,3-benzedicarboxylate(-2) ligand. In all these complexes, the 2paoH molecule behaves as a N,N′-bidentate chelating (1.011) ligand. The 4paoH Cd(II) complexes reported are {[Cd(mal)(4paoH)(H2O)]}n [56], {[Cd(adi)(4paoH)2]·DMF}n [56], {[Cd(SO4)(4paoH)2(H2O)2]}n [57], [Cd2(O2CMe)4(4paoH)4]·4H2O [58], [Cd(O2CMe)2(4paoH)3]·3H2O [58], {[Cd(1,3-bdc)(4paoH)(H2O)2]·DMF·H2O}n [60], {[Cd(1,4-bdc)(4paoH)2(H2O)]·DMF}n [60] and {[Cd(1,4-bdc)(4paoH)2]·DMF}n [60], where mal2− is the malonate(-2) ligand and adi2− is the adipate(-2) ligand. As in 3, in all of the just mentioned complexes, 4paoH behaves as an N(pyridyl) [1.010] monodentate ligand. It seems that this monodentate coordination mode is the preferable one for Cd(II).

3. Experimental Section

3.1. Materials and Spectrocopic- Physical Measurements

Experimental manipulations were carried out under aerobic conditions. Deionized water was received from the in-house facility. Solvents and reagents were purchased from Sigma-Aldrich (Tanfrichen, Germany) and Alfa Aesar (Karlsruhe, Germany), and used as received without extra purification. The free ligands 2,6-diacetylpyridine dioxime (dapdoH2, Figure 1) and 2,6-pyridyl-diamidoxime (LH4, Figure 1) were synthesized by following the procedures published in the literature [61,62]. The products were recrystallized from refluxing EtOH, and their yields were >70%. The purity of the free organic ligands was checked by microanalyses and 1H NMR spectroscopy.
Carbon, hydrogen and nitrogen microanalyses were performed by the Instrumental Analysis Center of the University of Patras. FT-IR spectra were recorded using a Perkin-Elmer spectrometer, model 16PC, manufactured by Perkin-Elmer (Waltham, MA, USA); the samples were in the form of KBr pellets prepared under pressure. FT-Raman spectra were obtained in an EQUINOX spectrometer to which a Bruker (D) FRA—106/S component had been attached (Bruker, Karlsruhe, Germany); an R510 diode-pumped Nd:YAG laser at 1064 nm was used for Raman excitation with a maximum laser power of 500 mW on the sample, utilizing an average of 100 scans at 4 cm−1 resolution. 1H and 13C NMR spectra were recorded on a Bruker Avance DPX spectrometer (Bruker AVANCE, Billerica, MA, USA) at resonance frequencies of 400.13 MHz (1H) and 100.62 MHz (13C); the signals of the solvent (d6-DMSO) were used as a reference. Conductivity measurements were performed at room temperature (23–25 °C) in DMSO with a Metrohm-Herisau E-527 bridge (Herisau, Switzerland) and a cell of standard constant; the concentration of the solution was 10−3 M.

3.2. Preparation of the Complexes

A variety of reaction systems involving various anions of the Cd(II) sources, and different solvent media, reagent ratios, crystallization techniques, reaction times and temperatures were employed before finding the optimized conditions described below.
[CdI2(2paoH)2] (1): A solution of CdI2 (0.055 g, 0.20 mmol) in MeCN (1 mL) was added to a solution of 2paoH (0.049 g, 0.40 mmol) in CH2Cl2 (3 mL). The resulting colorless solution was stirred for 10 min, filtered and was allowed to slowly evaporate at room temperature. X-ray quality, colorless crystals of the product were obtained within 3 d. The crystals were collected by filtration, washed with cold MeOH (2 × 0.5 mL) and Et2O (2 × 2 mL), and dried in vacuo over anhydrous CaCl2. Yield: 74%. Anal. Calcd. (%) for C12H12N4CdI2O2: C, 23.61; H, 1.99; N, 9.18. Found (%): C, 23.24; H, 1.91; N, 9.15. IR (KBr, cm1): 3324 w, 3314 w, 3304 w, 3228 wb, 3090 w, 3060 w, 1640 m, 1594 s, 1480 s, 1442 s, 1480 s, 1442 s, 1408 s, 1306 m, 1290 s, 1282 sh, 1256 s, 1212 m, 1150 m, 1106 m, 1052 w, 1004 s, 960 m, 932 s, 886 s, 774 s, 742 m, 672 s, 646 m, 600 m, 568 m, 510 s, 466 w. Raman (cm1): 3054 m, 3003 w, 1562 m, 1642 s, 1632 m, 1611 m, 1570 s, 1550 s, 1498 m, 1468 m, 1427 w, 1406 m, 1304 w, 1243 w, 1222 s, 1212 s, 1171 w, 1099 m, 1048 w, 1017 m, 925 w, 895 w, 772 w, 670 w, 506 w, 403 m, 301 m, 219 m, 137 s, 127 s, 117 s, 97 m, 76 m, 66 w. 1H NMR (d6-DMSO, δ/ppm): 10.97 (s, 2H), 7.91 (d, 2H), 7.42 (s, 2H), 7.15 (mt, 4H), 6.72 (t, 2H). 13C NMR (d6-DMSO, δ/ppm): 173.1, 152.3, 149.6, 137.4, 124.6, 120.5. ΛΜ (DMSO, 10−3 M, 25 °C) = 4 S cm2 mol−1.
{[CdI2(3paoH)2]}n (2): A solution of CdI2 (0.110 g, 0.40 mmol) in EtOH (1 mL) was slowly added to a slightly warm (~40 °C) solution of 3paoH (0.098 g, 0.80 mmol) in the same solvent (3 mL). The resulting colorless solution was stirred for 15 min and stored in a closed vial at 5 °C. X-ray quality, colorless crystals of the product were precipitated within 24 h. The crystals were collected by filtration, washed with cold EtOH (0.5 mL) and Et2O (2 × 1 mL), and dried in air. The average yield was 65%. Anal. Calcd. (%) for C12H12N4CdI2O2: C, 23.61; H, 1.99; N, 9.18. Found (%): C, 23.90; H, 1.97; N, 9.35. IR (KBr, cm−1): 3488 mb, 1624 w, 1596 w, 1574 w, 1482 m, 1424 m, 1386 w, 1330 w, 1252 s, 1228 sh, 1186 m, 1118 w, 1090 w, 1050 w, 952 s, 936 sh, 882 s, 804 m, 690 s, 642 m, 534 m, 472 mb. Raman (cm−1): 3067 m, 1642 s, 1597 m, 1578 m, 1391 w, 1331 w, 1264 w, 1233 w, 1185 w, 1050 w, 1032 m, 884 w, 643 w, 380 w, 292 w, 245 sh. 1H NMR (d6-DMSO, δ/ppm): 11.57 (sb, 2H), 8.75 (dd, 2H), 8.57 (dd, 2H), 8.21 (s, 2H), 8.00 (mt, 2H), 7.45 (dd, 2H). ΛΜ (DMSO, 10−3M, 25 °C) = 6 S cm2 mol−1.
{[CdI2(4paoH)2]}n (3): A solution of CdI2 (0.055 g, 0.20 mmol) in Me2CO (1 mL) was slowly added to a solution of 4paoH (0.049 g, 0.40 mmol) in the same solvent (3 mL). The resulting colorless solution was stirred and stored at −10 °C for 1 month. No solid was noticed and the solution was layered with Et2O (2 mL) and allowed to stand undisturbed at room temperature. Slow mixing gave crystals suitable for single-crystal, X-ray crystallography within 12 d. The crystals were collected by filtration, washed with cold EtOH (0.5 mL) and Et2O (2 × 1 mL), and dried in air overnight. Yield: 78%. Anal. Calcd. (%) for C12H12N4CdI2O2: C, 23.61; H, 1.99; N, 9.18. Found (%): C, 23.87; H, 1.90; N, 9.07. IR (KBr, cm−1): 3420 sb, 3010 w, 1638 wb, 1610 m, 1505 w, 1498 sh, 1420 m, 1398 m, 1264 s, 1224 sh, 1060 wb, 1010 w, 962 s, 936 m, 888 w, 814 m, 658 m, 564 w, 510 s, 486 m, 404 m. Raman (cm1): 3071 m, 3059 w, 1622 sh, 1613 s, 1543 w, 1399 w, 1327 w, 1269 w, 1240 w, 1225 w, 1206 m, 1013 m, 891 w, 668 w, 291 w, 239 w. 1H NMR (d6-DMSO, δ/ppm): 11.82 (s, 2H), 8.60 (dd, 4H), 8.18 (s, 2H), 7.55 (dd, 4H). ΛM (DMSO, 10−3 M, 25 °C) = 5 S cm2 mol−1.
[CdI2(dapdoH2)]·2EtOH (4·2EtOH): A solution of CdI2 (0.055 g, 0.20 mmol) in EtOH (1 mL) was added to a slurry of dapdoH2 (0.039 g, 0.20 mmol) in the same solvent (6 mL). The resulting pale yellow suspension was filtered to remove a very small quantity of the ligand. The filtrate was stirred for 2–3 min and stored in a closed vial at room temperature. X-ray quality, colorless crystals of the product were formed within 24 h which were collected by filtration, washed with cold EtOH (0.5 mL) and Et2O (2 × 1 mL), and dried in vacuo over anhydrous CaCl2. Yield: 28%. The sample was analyzed without lattice EtOH molecules. Anal. Calcd. (%) for C9H11N3CdI2O2: C, 19.32; H, 1.99; N, 7.51. Found (%): C, 19.35; H, 2.08; N, 8.11 IR (KBr, cm−1): 3422 s, 3178 mb, 3084 w, 3036 w, 2922 w, 1594 m, 1472 m, 1364 m, 1316 w, 1292 m, 1264 m, 1194 w, 1156 w, 1130 w, 1038 s, 950 m, 830 w, 808 m, 732w, 690 m, 654 w, 554 m, 508 w, 428 w. Raman (cm−1): 3076 m, 3065 sh, 2932 m, 1638 m, 1589sh, 1564 s, 1476 sh, 1464 w, 1429 w, 1358 m, 1314 w, 1262 w, 1129 w, 1015 s, 832 w, 760 w, 656 w, 448 w. 1H NMR (d6- DMSO, δ/ppm): 11.51 (s, 2H), 7.80 (mt, 3H), 2.25 (s, 6H). ΛΜ (DMSO, 10−3 M, 25 °C) = 8 S cm2 mol−1.
[CdI2(L′H2)] (5): A solution of CdI2 (0.055 g, 0.20 mmol) in EtOH (1 mL) was added to a solution of LH4 (0.039 g, 0.20 mmol) in the same solvent (7 mL). The resulting suspension was filtered to remove a small quantity of a precipitated material (5a). The filtrate was stirred for 2–3 min and stored in a closed vial at room temperature for 4d. The color of the reaction solution (filtrate) turned gradually to pink-pale violet during this time. This pink solution was placed at 5 °C, and X-ray quality, colorless crystals (from a pink solution!) were grown within a period of 7–8 d. The crystals were collected by filtration, washed with cold EtOH (0.5 mL) and Et2O (2 × 1 mL), and dried in air overnight. Yield: 41%. The sample 5a was analyzed as [CdI2(LH4)]. Anal. Calcd. (%) for C7H9N5CdI2O2: C, 14.97; H, 1.62; N, 12.48. Found (%): C, 14.60; H, 1.67; N, 12.63. IR spectrum (KBr, cm−1) for 5a: 3480 s, 3432 sb, 3372 sb, 3070 wb, 1654 s, 1604 sh, 1582 s, 1484 m, 1402 m, 1340 s, 1214 w, 1155 w, 1105 w, 1076 s, 1018 m, 986 s, 902 m, 810 s, 732 m, 696 s, 656 m, 522 mb, 460 w, 446 w. 1H NMR spectrum (d6-DMSO, δ/ppm) for 5a: 9.85 (s, 2H), 7.81 (mt, 3H), 6.32 (sb, 4H). Analytical data for the product 5: Anal. Calcd. (%) for C11H15N3CdI2O2: C, 22.49; H, 2.58; N, 7.15. Found (%): C, 28.93; H, 2.71; N, 7.07. IR spectrum (KBr, cm−1) for 5: 3444 mb, 3292 mb, 2984 m, 1648 m, 1578 s, 1458 m, 1404 m, 1378 m, 1344 s, 1290 m, 1198 m, 1146 s, 1106 m, 1018 m, 930 w, 838 m, 796 m, 748 m, 696 w, 660 m, 434 w. The ΛΜ values of both 5 and 5a in DMSO are ~5 S cm2 mol−1.

3.3. Single-Crystal X-ray Crystallography

Colorless crystals of 1, 2, 3, 4·2EtOH and 5 were taken from the mother liquor and immediately cooled to 160 (1, 3, 4·2EtOH, 5) or 170 (2) K. Diffraction data were collected on a Rigaku R-AXIS Image Plate diffractometer (Rigaku Americas Corporation, The Woodlands, TX, USA; European Department at Karlsruhe, Germany) using graphite-monochromated Cu Ka (1, 2) or Mo Ka (3, 4·2EtOH, 5) radiations. Data collection (ω-scans) and processing (cell refinement, data reduction, and empirical/numerical absorption correction) were performed using the CrystalClear program package [63]. The structures were solved by direct methods using SHELΧS, ver. 2013/1 [64] and refined by full-matrix least-squares techniques on F2 with SHELXL, ver. 2014/6 [65]. Most H atoms were introduced at calculated positions as riding on their corresponding bonded atoms. All non-H atoms were refined anisotropically. Plots of the structures were drawn using the Diamond 3 program package [66].
Crystallographic data have been deposited with the Cambridge Crystallographic Data Center, Nos 2142693 (1), 2142694 (2), 2142695 (3), 2142696 (4·2EtOH) and 2142697 (5). Copies of the data can be obtained free of charge upon application to CCDC, 12 Union Road, Cambridge, CB2 1EZ, UK: Tel.: +(44)-1223-762910; Fax: +(44)-1223-336033; E-mail: [email protected].

4. Conclusions in Brief

According to our opinion, the important chemical messages of this work are: (a) The reported complexes enrich the coordination chemistry of 2-pyridyl oximes, especially that with Cd(II). (b) The molecular structures and supramolecular features of the complexes are interesting; and (c) The interesting Cd(II)-assisted/promoted LH4 → L’H2 transformation has been observed for the first time and this is a new, welcome example in the area of the reactivity of coordinated amidoximes.
From the viewpoint of the solvent extraction of toxic Cd(II) using 2-pyridyl oximes as extractants (which stimulated the present efforts) our inorganic chemistry approach has firmly confirmed the molecular basis of the excellent extraction ability of 2PC12 and 2PC14, and the poor one for 4PC12, 4PC14. With the drawbacks mentioned in Introduction, the very good extraction ability of the 2-pyridyl ketoximes can be attributed to the chelating nature of the extractants as structurally proven by the 2paoH-containing complex 1; this chelating behavior results in thermodynamically stable complexes of Cd(II) with the extractant, favoring this process. The monodentate coordination of 4-pyridyl ketoximes (as structurally proven in the 4paoH-containg compound 3) seems to be responsible for the poor performance of 4PC12 and 4PC14. In an analogous manner (as proven in the 3paoH-conataining compound 2), extractants such as 3PC12 and 3PC14 (i.e., those containing the oxime group at position 3 of the pyridyl ring; Figure 1) are predicted to disfavor the extraction process; such “real” experiments are not available.
We tentatively propose that the structurally established tridentate chelating character of dapdoH2 towards Cd(II) in complex 4·2EtOH and the presumable such behavior of LH4 in complex 5a could be exploited to develop extractants consisting of one pyridyl ring and two oxime groups that contain long alkyl components at the 2- and 6-positions, or even to carry out experiments with the bis(amidoxime) compound LH4 (or better with derivatives containing aliphatic substituents on the pyridyl ring to ensure good solubility in non-polar organic solvents). The decomposition of 4 in DMSO (as evidenced by its 1H NMR spectrum in d6-DMSO, see Section 2.2) should not be disappointing, since this solvent is an efficient donor forming complexes with Cd(II) and favoring decomposition; the presence of coordinated dapdoH2-containing species in CDCl3 (albeit evidenced by poor quality 1H NMR spectra) is a good sign towards the use of 2,6-pyridyl dioximes as extractants for toxic Cd(II). The scientific community dealing with solvent extraction experiments might obtain good results working with tridentate dioximes.

Supplementary Materials

The following supporting information can be downloaded: Figure S1: The FT-IR spectrum (KBr/cm−1) of free paoH; Figure S2: The FT-Raman spectrum of free paoH; Figure S3: The FT-Raman spectrum of {[CdI2(4paoH)2]}n (3); Figure S4: The FT-IR spectrum (KBr/cm−1) of compound [CdI2(LH4)] (5a); Figure S5: The FT-IR spectrum (KBr/cm1) of complex [CdI2(L’H)2] (5); Figure S6: The 1H NMR spectrum (δ/ppm) of the free ligand 2paoH in d6-DMSO; Figure S7: The 1H NMR spectrum (δ/ppm) of complex {[CdI2(3paoH)]}n (2)x in d6-DMSO in the 9.0–7.33 ppm region; Figure S8: The 1H NMR spectrum (δ/ppm) of compound {[CdI2(4paoH)2]}n (3) in d6-DMSO in the 12.0–7.3 ppm region; Figure S9: A portion of one chain, parallel to the crystallographic direction [101], in the crystal structure of [CdI2(2paoH)2] (1). The thick dashed orange and green lines represent the intramolecular O1-H(O1)I1’ and O1’-H(O1’)I1 hydrogen bonds [(‘)= −x, y, −z + 1/2] and the π-π interactions, respectively; Figure S10: Layers of chains parallel to the (110) plane in the crystal structure of {[CdI2(3paoH)2]}n (2). The layers are formed through O-HI hydrogen bonds (thick dashed cyan lines); see text for more details; Figure S11: Layers of chains parallel to the (110) plane in the crystal structure of {[CdI2(4paoH)2]}n (3). The layers are formed through O-HI hydrogen bonds (thick dashed cyan lines); see text for the metric parameters; Figure S12: Formation of chains parallel to the a axis and layers parallel to the (001) crystallographic plane in the crystal structure of complex [CdI2(L’H2)] (5). The dashed thick orange line and the cyan lines represent the C(methyl)-Hπ and C(methyl)-HI interactions; see Table S2 for metric parameters; Table S1: Hydrogen bonding interactions (Å, deg) in the crystal structure of [CdI2(dapdoH2)]·2EtOH (2EtOH); Table S2: Hydrogen bonding interactions (Å, deg) in the crystal structure of [CdI2(L’H2)] (5).

Author Contributions

A.R., Z.G.L. and V.A. contributed towards the synthesis, crystallization, and conventional characterization of the complexes. Z.G.L. also contributed to the interpretation of the results and performed the Raman experiments. C.P.R. and V.P. collected single-crystal X-ray crystallographic data, solved the structures, and performed their refinements; the latter also investigated the supramolecular characteristics of the crystal structures and wrote the relevant part of the paper. K.F.K., C.T.C. and S.P.P. coordinated the research and wrote the paper based on the reports of their collaborators. S.P.P. coordinated the cooperation between the teams and submitted the manuscript. All the authors exchanged opinions concerning the progress of the experiments and commented on the various drafts of the paper. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Data Availability Statement

Not applicable.

Acknowledgments

The authors thank the MSc program “Analytical Chemistry and Nanotechnology” of the Chemistry Department of the University of Patras (Patras, Greece) for providing us with funding for the purchase of organic solvents. The authors also thank the head of Laboratory of Applied Molecular Spectroscopy (George A. Voyiatzis) at ICE-HT/FORTH for the access to the Raman setup.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Wu, T.; Liu, C.; Kong, B.; Sun, J.; Gorig, Y.; Liu, K.; ** the coordination chemistry of CdII using 113Cd solid-state NMR. Chem. Commun. 2016, 52, 10680–10683. [Google Scholar] [CrossRef] [PubMed]
  2. Manahan, S.E. Environmental Chemistry, 6th ed.; Lewis Publishers: Boca Raton, FL, USA, 1994. [Google Scholar]
  3. Watson, J.S. Separation Methods for Waste and Environmental Applications; Marcel Dekker: New York, NY, USA, 1999. [Google Scholar]
  4. Kefalas, E.T.; Dakanali, M.; Panagiotidis, P.; Raptopoulou, C.P.; Terzis, A.; Mavromoustakos, T.; Kyrikou, I.; Karligiano, N.; Bino, A.; Salifoglou, A. pH-specific aqueous synthetic chemistry in the binary cadmium(II)–citrate system. Gaining insight into cadmium(II)–citrate speciation with relevance to cadmium toxixity. Inorg. Chem. 2005, 44, 4818–4828. [Google Scholar] [CrossRef] [PubMed]
  5. Montero-Jiménez, M.; Fernández, L.; Alvarado, J.; Criollo, M.; Jadán, M.; Chuquer, D.; Espinoza-Montero, P. Evaluation of the cadmium accumulation in Tamarillo cells (Solanum betaceum) by indirect electrochemical detection of cysteine-rich peptides. Molecules 2019, 24, 2196. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  6. Shiraishi, T.; Tamada, M.; Saito, K.; Sugo, T. Recovery of cadmium from waste of scallop processing with amidoxime adsorbent synthesized by graft-polymerization. Radiat. Phys. Chem. 2003, 66, 43–47. [Google Scholar] [CrossRef]
  7. Alizadeh, T.; Sharifi, A.R.; Canjali, M.R. A new bio-compatible Cd2+-selective nanostructured fluorescent imprinted polymer for cadmium ion sensing in aqueous media and its application in bio imaging in Vero cells. RSC Adv. 2020, 10, 4110–4117. [Google Scholar] [CrossRef] [Green Version]
  8. Waalkes, M.P. Cadmium carcinogenesis. Mutat. Res. 2003, 533, 107–120. [Google Scholar] [CrossRef] [PubMed]
  9. Tomaszewska, M.; Borowiak-Resterna, A.; Olszanowski, A. Cadmium extraction from chloride solutions with model N-alkyl and N,N-dialkyl-pyridine-carboxamides. Hydrometallurgy 2007, 85, 116–126. [Google Scholar] [CrossRef]
  10. Parus, A.; Wieszczycka, K.; Olszanowski, A. Solvent extraction of cadmium(II) from chloride solutions by pyridyl ketoximes. Hydrometallurgy 2011, 105, 284–289. [Google Scholar] [CrossRef]
  11. Klonowska-Wieszczycka, K.; Olszanowski, A.; Parus, A.; Zydorczak, B. Removal of copper(II) from chloride solutions using hydrophobic pyridyl ketone oximes. Solvent Extr. Ion Exch. 2009, 27, 50–62. [Google Scholar] [CrossRef]
  12. Efthymiou, C.G.; Cunha-Silva, L.; Perlepes, S.P.; Brechin, E.K.; Inglis, R.; Evangelisti, M.; Papatriantafyllopoulou, C. In search of molecules displaying ferromagnetic exchange: Multiple-decker Ni12 and Ni16 complexes from the use of pyridine-2-amidoxime. Dalton Trans. 2016, 45, 17409–17419. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  13. Zahra, M.; Zulfigar, S.; Skene, W.G.; Sarwar, M.I. Efficient uptake of Cd(II) and Pb(II) ions by aromatic polyamidoximes. Ind. Eng. Chem. Res. 2018, 57, 15243–15253. [Google Scholar] [CrossRef]
  14. Milios, C.J.; Stamatatos, T.C.; Perlepes, S.P. The coordination chemistry of pyridyl oximes. Polyhedron 2006, 25, 134–194. [Google Scholar] [CrossRef]
  15. Konidaris, K.F.; Polyzou, C.D.; Kostakis, G.E.; Tasiopoulos, A.J.; Roubeau, O.; Teat, S.J.; Manessi-Zoupa, E.; Powell, A.K.; Perlepes, S.P. Metal ion-assisted transformations of 2-pyridinealdoxime and hexafluorophosphate. Dalton Trans. 2012, 41, 2862–2865. [Google Scholar] [CrossRef]
  16. Tsantis, S.T.; Zagoraiou, E.; Savvidou, A.; Raptopoulou, C.P.; Psycharis, V.; Holynska, M.; Perlepes, S.P. Binding of oxime group to uranyl ion. Dalton Trans. 2016, 45, 9307–9319. [Google Scholar] [CrossRef]
  17. Polyzou, C.D.; Koumousi, E.S.; Lada, Z.G.; Raptopoulou, C.P.; Psycharis, V.; Rouzières, M.; Tsipis, A.C.; Mathonière, C.; Clérac, R.; Perlepes, S.P. “Switching on” the single-molecule magnet properties within a series of dinuclear cobalt(III)-dysposium(III) 2-pyridyloximate complexes. Dalton Trans. 2017, 46, 14812–14825. [Google Scholar] [CrossRef]
  18. Papatriantafyllopoulou, C.; Stamatatos, T.C.; Wernsdorfer, W.; Teat, S.J.; Tasiopoulos, A.J.; Escuer, A.; Perlepes, S.P. Combining azide, carboxylate and 2-pyridyloximate ligands in transition-metal chemistry: Ferromagnetic NiII5 clusters with a bowtie skeleton. Inorg. Chem. 2010, 49, 10486–10496. [Google Scholar] [CrossRef]
  19. Nikolaou, H.; Terzis, A.; Raptopoulou, C.P.; Psycharis, V.; Bekiari, V.; Perlepes, S.P. Unique dinuclear, tetrakis (nitrato–O,O′)-bridged lanthanide(III) complexes from the use of pyridine-2-amidoxime: Synthesis, structural studies and spectroscopic characterization. J. Surf. Interfaces Mater. 2014, 2, 311–318. [Google Scholar] [CrossRef]
  20. Anastasiadis, N.C.; Polyzou, C.D.; Kostakis, G.E.; Bekiari, V.; Lan, Y.; Perlepes, S.P.; Konidaris, K.F.; Powell, A.K. Dinuclear lanthanide(III)/zinc(II) complexes with methyl 2-pyridyl ketone oxime. Dalton Trans. 2015, 44, 19791–19795. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  21. Stamatatos, T.C.; Foguet-Albiol, D.; Lee, S.-C.; Stoumpos, C.C.; Raptopoulou, C.P.; Terzis, A.; Wernsdorfer, W.; Hill, S.O.; Perlepes, S.P.; Christou, G. “Switching on” the properties of single-molecule magnetism in triangular manganese(III) complexes. J. Am. Chem. Soc. 2007, 129, 9484–9499. [Google Scholar] [CrossRef]
  22. Polyzou, C.D.; Lada, Z.G.; Terzis, A.; Raptopoulou, C.P.; Psycharis, V.; Perlepes, S.P. The fac diastereoisomer of tris (2-pyridinealdoximato)cobalt(III) and a cationic cobalt(III) complex containing both the neutral and and anionic forms of the ligand: Synthetic, structural and spectroscopic studies. Polyhedron 2014, 79, 29–36. [Google Scholar] [CrossRef]
  23. Stamatatos, T.C.; Escuer, A.; Abboud, K.A.; Raptopoulou, C.P.; Perlepes, S.P.; Christou, G. Unusual structural types in nickel cluster chemistry from the use of pyridyl oximes: Ni5, Ni12Na2, and Ni14 clusters. Inorg. Chem. 2008, 47, 11825–11838. [Google Scholar] [CrossRef]
  24. Polyzou, C.D.; Nikolaou, H.; Raptopoulou, C.P.; Konidaris, K.F.; Bekiari, V.; Psycharis, V.; Perlepes, S.P. Dinuclear lanthanide(III) complexes from the use of methyl 2-pyridyl ketoxime: Synthetic, structural and physical studies. Molecules 2021, 26, 1622. [Google Scholar] [CrossRef]
  25. Papatriantafyllopoulou, C.; Kostakis, G.E.; Raptopoulou, C.P.; Terzis, A.; Perlepes, S.P.; Plakatouras, J.C. Investigation of the MSO4·xH2O (M = Zn, x = 7; M = Cd, x = 8/3)/methyl 2-pyridyl ketone oxime reaction system: A novel Cd(II) coordination polymer versus mononuclear and dinuclear Zn(II) complexes. Inorg. Chim. Acta 2009, 362, 2361–2370. [Google Scholar] [CrossRef]
  26. Stamatatos, T.C.; Katsoulakou, E.; Nastopoulos, V.; Raptopoulou, C.P.; Manessi-Zoupa, E.; Perlepes, S.P. Cadmium carboxylate chemistry: Preparation, crystal structure, and thermal and spectroscopic characterization of the one-dimensional polymer [Cd(O2CMe)(O2CPh)(H2O)2]n. Z. Nat. B 2003, 58, 1045–1054. [Google Scholar] [CrossRef]
  27. Papatriantafyllopoulou, C.; Raptopoulou, C.P.; Terzis, A.; Janssens, J.F.; Manessi-Zoupa, E.; Perlepes, S.P.; Plakatouras, J.C. Assembly of a helical zinc(II) chain and a two-dimensional cadmium(II) coordination polymer using picolinate and sulfate anions as bridging ligands. Polyhedron 2007, 26, 4053–4064. [Google Scholar] [CrossRef]
  28. Katsoulakou, E.; Konidaris, K.F.; Raptopoulou, C.P.; Psycharis, V.; Manessi-Zoupa, E.; Perlepes, S.P. Synthesis, X-ray structure and characterization of catena-bis (benzoate) bis {N,N-bis(2-hydroxyethyl)-glycinate}cadmium(II). Bioinorg. Chem. Applic. 2010, 2010, 281932. [Google Scholar]
  29. Katsoulakou, E.; Konidaris, K.F.; Terzis, A.; Raptopoulou, C.P.; Perlepes, S.P.; Manessi-Zoupa, E.; Kostakis, G.E. One-dimensional cadmium(II)/bicinate(−1) complexes: The role of the alkali metal ion used in the reaction medium. Polyhedron 2011, 30, 397–404. [Google Scholar] [CrossRef]
  30. Katsoulakou, E.; Bekiari, V.; Raptopoulou, C.P.; Terzis, A.; Manessi-Zoupa, E.; Powell, A.; Perlepes, S.P. Simultaneous coordination of a ketone by two cadmium(II) ions and conversion to its gem-diolate(−1) form. Inorg. Chem. Commun. 2011, 14, 1057–1060. [Google Scholar] [CrossRef]
  31. Bolotin, D.S.; Bokach, N.A.; Kukushkin, V.Y. Coordination chemistry and metal-involving reactions of amidoximes: Relevance to the chemistry of oximes and oxime ligands. Coord. Chem. Rev. 2016, 313, 62–93. [Google Scholar] [CrossRef] [Green Version]
  32. Bellamy, L.J. The Infrared Spectra of Complex Molecules; Methyen: London, UK, 1966; p. 249. [Google Scholar]
  33. Dolish, F.R.; Fateley, W.G.; Bentley, F.F. Characteristic Raman Frequencies of Organic Compounds; John Wiley and Sons: New York, NY, USA, 1974; pp. 135–137. [Google Scholar]
  34. Suvitha, A.; Periandy, S.; Boomadevi, S.; Govindarajan, M. Vibrational frequency analysis, FT-IR, FT-Raman, ab initio, HF and DFT studies, NBO, HOMO-LUMO and electronic structure calculations on pycolinaldehyde oxime. Spectrochim. Acta Part A 2014, 117, 216–224. [Google Scholar] [CrossRef]
  35. Lopez-Garriga, J.J.; Babcock, G.T.; Harrison, J.F. Factors influencing the C=N stretching frequency in neutral and protonated Schiff’s bases. J. Am. Chem. Soc. 1986, 108, 7241–7251. [Google Scholar] [CrossRef]
  36. Nordquest, K.W.; Phelps, D.W.; Little, W.F.; Hodgson, D.J. Metal-metal interactions in linear chains. The structure and characterization of bis (pyridine-2-carboxaldoximinato)platinum(II) dihydrate. J. Am. Chem. Soc. 1976, 98, 1104–1107. [Google Scholar] [CrossRef]
  37. Geary, W.J. The use of conductivity measurements in organic solvents for the characterization of coordination compounds. Coord. Chem. Rev. 1971, 7, 81–122. [Google Scholar] [CrossRef]
  38. Nieuwenhuyzen, M.; Wen, H.; Wilkins, C.J. Cadmium iodide complexes with dimethylsulphoxide, and their crystal structures. Z. Anorg. Allg. Chem. 1992, 615, 143–148. [Google Scholar] [CrossRef]
  39. Coxall, R.A.; Harris, S.G.; Henderson, D.K.; Parsons, S.; Tasker, P.A.; Winpenny, R.E.P. Inter-ligand reactions: In situ formation of new polydentate ligands. J. Chem. Soc. Dalton Trans. 2000, 2349–2356. [Google Scholar] [CrossRef]
  40. Shirvan, S.A.; Dezfuli, S.H. Bis(acetate-κO)bis(2-pyridinealdoxime-κ2N,N′)cadmium. Acta Cryst. E. 2012, 68, m1080–m1081. [Google Scholar] [CrossRef]
  41. Croitor, L.; Coropceanu, E.B.; Masunov, A.E.; Rivera-Jacquez, H.J.; Siminel, A.V.; Zelentsov, V.I.; Datsko, T.Y.; Fonari, M.S. Polymeric luminescent Zn(II) and Cd(II) dicarboxylates decorated by oxime ligands: Tuning the dimensionality and adsorption capacity. Cryst. Growth Des. 2014, 14, 3935–3948. [Google Scholar] [CrossRef]
  42. Croitor, L.; Coropseanu, E.B.; Masunov, A.E.; Rivera-Jacquez, H.J.; Siminel, A.V.; Fonari, M.S. Mechanism of nonlinear optical enhancement and supramolecular isomerism in 1D polymeric Zn(II) and Cd(II) sulfates with pyridine-4-aldoxime ligands. J. Phys. Chem. C 2014, 118, 9217–9227. [Google Scholar] [CrossRef]
  43. Coropseanu, E.B.; Croitor, L.; Siminel, A.V.; Fonari, M.S. Preparation, structural characterization and luminescence studies of mono- and binuclear Zn(II) and Cd(II) acetates with pyridine-4-aldoxime and pyridine-4-amidoxime ligands. Polyhedron 2014, 75, 73–80. [Google Scholar] [CrossRef]
  44. Croitor, L.; Coropceanu, E.B.; Siminel, A.V.; Masunov, A.E.; Fonari, M.S. From discrete molecules to one-dimensional coordination polymers containing Mn(II), Zn(II) or Cd(II) pyridine-2-aldoxime building unit. Polyhedron 2013, 60, 140–150. [Google Scholar] [CrossRef]
  45. Croitor, L.; Coropceanu, E.B.; Duca, G.; Siminel, A.V.; Fonari, M.S. Nine Mn(II), Zn(II) and Cd(II) mixed-ligand coordination networks with rigid dicarboxylate and pyridine-n-aldoxime ligands. Impact of the second ligand in the structures’ dimensionality and solvent capacity. Polyhedron 2017, 129, 9–21. [Google Scholar] [CrossRef]
  46. Bacuom, E.I.; Drago, R.S. Nickel(II) and nickel(IV) complexes of 2,6-diacetylpyridine dioxime. J. Am. Chem. Soc. 1971, 93, 6469–6475. [Google Scholar] [CrossRef]
  47. Pearse, G.A., Jr.; Bovenzi, B.A. The synthesis and crystal structures of pyridine-2,6-dicarboxamide oxime, C7H9N5O2, and its nickel(II) and copper(II) co-ordination compounds. J. Chem. Soc. Dalton Trans. 1997, 2793–2797. [Google Scholar] [CrossRef]
  48. CrystalClear; Rigaku: The Woodlands, TX, USA; MSC Inc.: The Woodlands, TX, USA, 2005.
  49. Sheldrick, G.M. A short history of SHELX. Acta Crystallogr. Sect. A 2008, 64, 112–122. [Google Scholar] [CrossRef] [Green Version]
  50. Sheldrick, G.M. Crystal structure refinement with SHELXL. Acta Crystallogr. Sect. C 2015, 71, 3–8. [Google Scholar] [CrossRef]
  51. Diamond, Crystal and Molecular Structure Visualization; Version 3.1; Crystal Impact: Bonn, Germany, 2018.
Figure 1. Structural formulae of the extractants and the ligands discussed in the text, and their abbreviations; the presence of four hydrogens in the abbreviation of 2,6-pyridyl diamidoxime (LH4) implies the known in the literature possibility of the single deprotonation of the -NH2 group [27].
Figure 1. Structural formulae of the extractants and the ligands discussed in the text, and their abbreviations; the presence of four hydrogens in the abbreviation of 2,6-pyridyl diamidoxime (LH4) implies the known in the literature possibility of the single deprotonation of the -NH2 group [27].
Molecules 27 01619 g001
Figure 2. Optimized experimental reaction conditions that lead to Cd(II) complexes (15); the lattice EtOH molecules in the dapdoH2 complex are not shown. The structural formula of the ligand L′H2 and a possible mechanism of the LH4 → L′H2 transformation are presented in Figure 3 (vide infra).
Figure 2. Optimized experimental reaction conditions that lead to Cd(II) complexes (15); the lattice EtOH molecules in the dapdoH2 complex are not shown. The structural formula of the ligand L′H2 and a possible mechanism of the LH4 → L′H2 transformation are presented in Figure 3 (vide infra).
Molecules 27 01619 g002
Figure 3. A simplified mechanistic proposal for the CdI2-promoted/assisted LH4 → L′H2 transformation. The I ions come from the metal source, while the H+ ions (as a matter of fact H3O+) from the first step of hydrolysis/solvolysis of [Cd(H2O)6]2+/[Cd(EtOH)6]2+; the water is contained in the organic solvent which is not anhydrous.
Figure 3. A simplified mechanistic proposal for the CdI2-promoted/assisted LH4 → L′H2 transformation. The I ions come from the metal source, while the H+ ions (as a matter of fact H3O+) from the first step of hydrolysis/solvolysis of [Cd(H2O)6]2+/[Cd(EtOH)6]2+; the water is contained in the organic solvent which is not anhydrous.
Molecules 27 01619 g003
Figure 4. The FT-Raman spectrum of the ligand 3paoH (black line) and the {[CdI2(3paoH)2]}n (2) complex (red line).
Figure 4. The FT-Raman spectrum of the ligand 3paoH (black line) and the {[CdI2(3paoH)2]}n (2) complex (red line).
Molecules 27 01619 g004
Figure 5. The FT-Raman spectrum of a well-dried (i.e., without lattice EtOH) sample of [CdI2(dapdoH2)] (4).
Figure 5. The FT-Raman spectrum of a well-dried (i.e., without lattice EtOH) sample of [CdI2(dapdoH2)] (4).
Molecules 27 01619 g005
Figure 6. The structure of the [CdI2(2paoH)2] molecule that is present in 1. The thermal ellipsoids are presented at the 50% probability level. Only the metal, the donor atoms and the oxime oxygen atoms are numbered. The dashed lines indicate intramolecular hydrogen bonds. Symmetry operation: (′) = x, y, −z + 1/2.
Figure 6. The structure of the [CdI2(2paoH)2] molecule that is present in 1. The thermal ellipsoids are presented at the 50% probability level. Only the metal, the donor atoms and the oxime oxygen atoms are numbered. The dashed lines indicate intramolecular hydrogen bonds. Symmetry operation: (′) = x, y, −z + 1/2.
Molecules 27 01619 g006
Figure 7. A small portion of the 1D chain that is present in the crystal structure of {[CdI2(3paoH)2]}n (2). The thermal ellipsoids are presented at the 50% probability level. Only the CdII atoms, the donor atoms and the N, O atoms of the oxime groups are numbered. Symmetry operations: (′)= −x, y, −z; (′′) = x, y − 1, z; (′′′) = −x, y − 1, −z; (*) = x, y + 1, z; (**) = −x, y + 1, −z.
Figure 7. A small portion of the 1D chain that is present in the crystal structure of {[CdI2(3paoH)2]}n (2). The thermal ellipsoids are presented at the 50% probability level. Only the CdII atoms, the donor atoms and the N, O atoms of the oxime groups are numbered. Symmetry operations: (′)= −x, y, −z; (′′) = x, y − 1, z; (′′′) = −x, y − 1, −z; (*) = x, y + 1, z; (**) = −x, y + 1, −z.
Molecules 27 01619 g007
Figure 8. A small portion of the 1D chain that is present in the crystal structure of {[CdI2(4paoH)2]}n (3). The thermal ellipsoids are presented at the 50% probability level. Only the CdII atoms, the donor atoms and the N, O atoms of the oxime groups are numbered. Symmetry operations: (′) = −x, y, −z; (′′) = −x, y − 1, −z; (′′′) = x, y − 1, z; (*) = x, y + 1, z; (**) = −x, y + 1, −z.
Figure 8. A small portion of the 1D chain that is present in the crystal structure of {[CdI2(4paoH)2]}n (3). The thermal ellipsoids are presented at the 50% probability level. Only the CdII atoms, the donor atoms and the N, O atoms of the oxime groups are numbered. Symmetry operations: (′) = −x, y, −z; (′′) = −x, y − 1, −z; (′′′) = x, y − 1, z; (*) = x, y + 1, z; (**) = −x, y + 1, −z.
Molecules 27 01619 g008
Figure 9. The molecules [CdI2(dapdoH2)] and EtOH that are present in the crystal structure of 4·2EtOH. The thermal ellipsoids are presented at the 50% probability level. The thick dashed orange lines indicate hydrogen bonds between the oxime groups and the lattice EtOH molecules.
Figure 9. The molecules [CdI2(dapdoH2)] and EtOH that are present in the crystal structure of 4·2EtOH. The thermal ellipsoids are presented at the 50% probability level. The thick dashed orange lines indicate hydrogen bonds between the oxime groups and the lattice EtOH molecules.
Molecules 27 01619 g009
Figure 10. Layers of molecules [CdI2(dapdoH2)] parallel to the (001) plane in the crystal structure of 4·2EtOH. The thick dashed dark and light green lines represent π-π interactions. The thick dashed cyan and brown lines represent the C4-H(C4)-I2 and C9-HC(C9)O2 hydrogen bonds, respectively. The hydrogen bonds involving the lattice EtOH molecules have not been drawn for clarity reasons; for more details see the above text and Table S1.
Figure 10. Layers of molecules [CdI2(dapdoH2)] parallel to the (001) plane in the crystal structure of 4·2EtOH. The thick dashed dark and light green lines represent π-π interactions. The thick dashed cyan and brown lines represent the C4-H(C4)-I2 and C9-HC(C9)O2 hydrogen bonds, respectively. The hydrogen bonds involving the lattice EtOH molecules have not been drawn for clarity reasons; for more details see the above text and Table S1.
Molecules 27 01619 g010
Figure 11. The structure of [CdI2(L′H2)] molecule that is present in 5. The thermal ellipsoids are presented at the 50% probability level. Symmetry operation: (′) = −x, −y + 3/2, z.
Figure 11. The structure of [CdI2(L′H2)] molecule that is present in 5. The thermal ellipsoids are presented at the 50% probability level. Symmetry operation: (′) = −x, −y + 3/2, z.
Molecules 27 01619 g011
Table 1. Crystallographic data and refinement parameters for the structures of 1, 2, 3, 2EtOH and 5.
Table 1. Crystallographic data and refinement parameters for the structures of 1, 2, 3, 2EtOH and 5.
Parameter[CdI2(2paoH)2] (1){[CdI2(3paoH)2]}n (2){[CdI2(4paoH)2]}n (3)[CdI2(dapdoH2)]·2EtOH (4·2EtOH)[CdI2(L′H2)] (5)
Empirical formulaC12H12CdI2N4O2C12H12CdI2N4O2C12H12CdI2N4O2C13H23CdI2N3O4C11H15CdI2N3O2
Formula weight610.46610.46610.46651.54587.46
Crystal systemmonoclinicMonoclinicmonoclinictriclinicorthorhombic
Space groupC2/cC2/mC2/mP 1 ¯ Pcnb
ColorcolorlessColorlesscolorlesscolorlesscolorless
Crystal size, mm0.41 × 0.20 × 0.170.21 × 0.21 × 0.060.23 × 0.13 × 0.050.35 × 0.18 × 0.130.35 × 0.19 × 0.08
a, Å7.9245(2)25.0595(11)24.8525(7)8.0032(19)6.4986(3)
b, Å13.7155(2)4.1616(2)4.1618(10)9.824(2)15.0095(6)
c, Å15.6259(2)7.9660(4)8.1913(3)14.215(3)17.7360(7)
α, ° 90.090.090.081.439(6)90.0
β, °101.35(1)98.413(1)98.223(1)78.146(5)90.0
γ, °90.090.090.074.771(6)90.0
Volume, Å31665.13(8)821.82(7)838.52(4)1050.0(4)1729.98(13)
Z42224
Temperature, K160170160160160
Radiation, Å/2θmaxCu Kα (1.54178)/130.0Cu Kα (1.54178)/129.8Mo Kα (0.71073)/54.0Mo Kα (0.71073)/54.0Mo Kα (0.71073)/54.0
Calculated density, g·cm−32.4352.4672.4182.0612.256
Absorption coefficient, mm−139.7140.234.994.004.83
Number of measured, independent, and observed [I > 2σ(I)] reflections10,738, 1385, 13335658, 718, 66110,187, 1042, 98832,407, 4564, 418935,910, 1875, 1753
Number of parameters98656521689
Final R indices [I > 2σ(Ι)] aR1 = 0.0384, wR2 = 0.0902R1 = 0.0499, wR2 = 0.1079R1 = 0.0166, wR2 = 0.0398R1 = 0.0224, wR2 = 0.0570R1 = 0.0337, wR2 = 0.0707
Goodness-of-fit on F21.061.101.051.021.11
Largest differences peak and hole (e Å−3)1.28/−1.661.22/−1.200.66/−0.520.53/−0.740.50/−0.84
a R1 = Σ(|Fo| − |Fc|)/Σ(|Fo|), wR2 = {Σ[w(Fo2Fc2)2]/Σ[w(Fo2)2]}1/2, w = 1⁄[σ2(Fo2) + (αP)2 + bP], where P = [max (Fo2,0) + 2Fc2]/3. (α = 0.0029 and b = 7.5036 for 1; α = 0.0519 and b = 2.3524 for 2; α = 0.0232 and b = 0.8669 for 3; α = 0.0281 and b = 1.0643 for 2EtOH; α = 0.0151, b = 7.8149 for 5).
Table 2. Selected bond lengths (Å) and angles (°) for the complex [CdI2(2paoH)2] (1) a.
Table 2. Selected bond lengths (Å) and angles (°) for the complex [CdI2(2paoH)2] (1) a.
Bond Lengths (Å)Bond Angles (°)
Cd1-I12.829(1)I1-Cd1-I1 ′103.4(1)
Cd1-N12.457(6)I1-Cd1-N1114.8(1)
Cd1-N22.402(5)I1-Cd1-N1 ′86.8(1)
C1-N11.253(8)I1-Cd1-N289.8(1)
N1-O11.398(7)I1-Cd1-N2 ′153.6(1)
N1-Cd1-N1 ′145.8(2)
N1-Cd1-N266.9(2)
N1-Cd1-N2 ′88.2(2)
N2-Cd1-N2 ′87.8(2)
a Symmetry code: (′) = −x, y, −z + 1/2. Atoms C1 and C1′, not labelled in Figure 6, are the oxime carbon atoms of coordinated 2paoH.
Table 3. Selected interatomic distances (Å) and angles (°) for the polymeric compound {[CdI2(3paoH)2]}n (2) a.
Table 3. Selected interatomic distances (Å) and angles (°) for the polymeric compound {[CdI2(3paoH)2]}n (2) a.
Interatomic Distances (Å)Interatomic Angles (°)
Cd1-N2/N2 ′2.353(9)N2-Cd1-N2 ′ = I1-Cd1-I1 ′′ = I1 ′-Cd1-I1 *180.0(1)
Cd1-I1/I1 ′′2.986(1)I1-Cd1-I1 ′ = I1 *-Cd1-I1 **91.6(1)
Cd1Cd1 ′′4.162(1)I1-Cd1-I1 * = I1 ′-Cd1-I1 **88.4(2)
N1-O1 = N1 ′-O1 ′ = N1 ′′-O1 ′′ = N1 ′′′-O1 ′′′1.423(14)I1 ′-Cd1-N2 = I1-Cd1-N2 ′ = I1 *-Cd1-N2 ′ = I1 **-Cd1-N290.9(2)
I1-Cd1-N2 = I1 ′-Cd1-N2 ′ = I1 *-Cd1-N2 = I1 ′′-Cd1-N2 ′89.1(2)
Cd1-I1-Cd1 ′′88.4(2)
a Symmetry codes: () = −x, y, −z; (′′) = x, y − 1, z; (′′′) = −x, y − 1, −z; (*) = x, y + 1, z; (**) = −x, y + 1, −z.
Table 4. Selected interatomic distances (Å) and angles (°) for the polymeric compound {[CdI2(4paoH)2]}n (3) a.
Table 4. Selected interatomic distances (Å) and angles (°) for the polymeric compound {[CdI2(4paoH)2]}n (3) a.
Interatomic Distances (Å)Interatomic Angles (°)
Cd1-N2/N2 ′2.353(2)N2-Cd1-N2 ′ = I1-Cd1-I1 ′′ = I1 ′-Cd1-I1 *180.0(1)
Cd1-I1/I1 ′/I1 */I1 **2.991(1)I1-Cd1-I1 ′ = I1 *-Cd1-I1 ′′91.8(1)
Cd1Cd1 ′′′4.162(1)I1-Cd1-I1 * = I1 ′-Cd1-I1 ′′88.2(1)
N1-O1=N1 ′-O1 ′1.403(4)I1-Cd1-N2 ′ = I1 ′-Cd1-N2 = I1 *-Cd1-N2 ′ = I1 **-Cd1-N290.7(1)
I1-Cd1-N2 = I1 ′-Cd1-N2 ′ = I1 *-Cd1-N2 = I1 **-Cd1-N2 ′89.3(1)
Cd1-I1-Cd1 ′′′88.2(1)
a Symmetry codes: () = −x, y, −z; (′′) = −x, y − 1, −z; (′′′) = x, y − 1, z; (*) = x, y + 1, z; (**) = −x, y + 1, −z.
Table 5. Selected bond lengths (Å) and angles (°) for complex [CdI2(dapdoH2)]·2EtOH (2EtOH).
Table 5. Selected bond lengths (Å) and angles (°) for complex [CdI2(dapdoH2)]·2EtOH (2EtOH).
Bond Lengths (Å)Bond Angles (°)
Cd-I12.722(1)I1-Cd-I2126.7(1)
Cd-I22.733(1)I1-Cd-N1100.2(1)
Cd-N12.421(2)I1-Cd-N2117.8(1)
Cd-N22.333(2)I1-Cd-N398.8(1)
Cd-N32.443(2)I2-Cd-N1101.0(1)
C2-N11.279(4)I2-Cd-N2115.4(1)
N1-O11.380 (3)I2-Cd-N399.6(1)
C8-N31.275(3)N1-Cd-N267.5(1)
N3-O21.384(3)N1-Cd-N3134.9(1)
N2-Cd-N367.5(1)
Table 6. Selected bond lengths (Å) and angles (°) for complex [CdI2(L′H2)] (5) a.
Table 6. Selected bond lengths (Å) and angles (°) for complex [CdI2(L′H2)] (5) a.
Bond Lengths (Å)Bond Angles (°)
Cd1-I12.738(1)I1-Cd1-I1 ′119.4(1)
Cd1-N12.423(4)I1-Cd1-N1100.7(1)
Cd1-N22.344(5)I1-Cd1-N2120.3(1)
C3-N11.268(6)I1-Cd1-N1 ′100.9(1)
C3-O11.335(6)N1-Cd1-N268.2(1)
O1-C21.451(6)N1-Cd1-N1 ′136.4(1)
C1-C21.489(7)
a Symmetry code: (′) = −x, −y + 3/2, z.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Routzomani, A.; Lada, Z.G.; Angelidou, V.; P. Raptopoulou, C.; Psycharis, V.; Konidaris, K.F.; Chasapis, C.T.; Perlepes, S.P. Confirming the Molecular Basis of the Solvent Extraction of Cadmium(II) Using 2-Pyridyl Oximes through a Synthetic Inorganic Chemistry Approach and a Proposal for More Efficient Extractants. Molecules 2022, 27, 1619. https://doi.org/10.3390/molecules27051619

AMA Style

Routzomani A, Lada ZG, Angelidou V, P. Raptopoulou C, Psycharis V, Konidaris KF, Chasapis CT, Perlepes SP. Confirming the Molecular Basis of the Solvent Extraction of Cadmium(II) Using 2-Pyridyl Oximes through a Synthetic Inorganic Chemistry Approach and a Proposal for More Efficient Extractants. Molecules. 2022; 27(5):1619. https://doi.org/10.3390/molecules27051619

Chicago/Turabian Style

Routzomani, Anastasia, Zoi G. Lada, Varvara Angelidou, Catherine P. Raptopoulou, Vassilis Psycharis, Konstantis F. Konidaris, Christos T. Chasapis, and Spyros P. Perlepes. 2022. "Confirming the Molecular Basis of the Solvent Extraction of Cadmium(II) Using 2-Pyridyl Oximes through a Synthetic Inorganic Chemistry Approach and a Proposal for More Efficient Extractants" Molecules 27, no. 5: 1619. https://doi.org/10.3390/molecules27051619

Article Metrics

Back to TopTop