Next Article in Journal
Intensity Ratio of / in Selected Elements from Mg to Cu, and the Chemical Effects of Cr 1,2 Diagram Lines and Cr / Intensity Ratio in Cr Compounds
Next Article in Special Issue
First 24-Membered Macrocyclic 1,10-Phenanthroline-2,9-Diamides—An Efficient Switch from Acidic to Alkaline Extraction of f-Elements
Previous Article in Journal
Co-Catalyzed Asymmetric Hydrogenation. The Same Enantioselection Pattern for Different Mechanisms
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Pyrrolidine-Derived Phenanthroline Diamides: An Influence of Fluorine Atoms on the Coordination of Lu(III) and Some Other f-Elements and Their Solvent Extraction

by
Nane A. Avagyan
1,
Pavel S. Lemport
1,
Mariia V. Evsiunina
1,
Petr I. Matveev
1,
Svetlana A. Aksenova
2,3,
Yulia V. Nelyubina
2,
Alexandr V. Yatsenko
1,
Viktor A. Tafeenko
1,
Vladimir G. Petrov
1,
Yuri A. Ustynyuk
1,
Xihe Bi
4 and
Valentine G. Nenajdenko
1,*
1
Department of Chemistry, Lomonosov Moscow State University, 119991 Moscow, Russia
2
A.N. Nesmeyanov Institute of Organoelement Compounds, Russian Academy of Sciences, 119334 Moscow, Russia
3
Moscow Institute of Physics and Technology, National Research University, Institutskiy per. 9, 141700 Dolgoprudny, Russia
4
Department of Chemistry, Northeast Normal University, 5286 Renmin Street, Changchun 130024, China
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2023, 24(6), 5569; https://doi.org/10.3390/ijms24065569
Submission received: 21 February 2023 / Revised: 10 March 2023 / Accepted: 13 March 2023 / Published: 14 March 2023
(This article belongs to the Special Issue Advanced Ligands for Coordination Chemistry)

Abstract

:
Three pyrrolidine-derived phenanthroline diamides were studied as ligands for lutetium trinitrate. The structural features of the complexes have been studied using various spectral methods and X-ray. The presence of halogen atoms in the structure of phenanthroline ligands has a significant impact on both the coordination number of lutetium and the number of solvate water molecules in the internal coordination sphere. The stability constants of complexes with La(NO3)3, Nd(NO3)3, Eu(NO3)3, and Lu(NO3)3 were measured to demonstrate higher efficiency of fluorinated ligands. NMR titration was performed for this ligand, and it was found that complexation with lutetium leads to an approximately 13 ppm shift of the corresponding signal in the 19F NMR spectrum. The possibility of formation of a polymeric oxo-complex of this ligand with lutetium nitrate was demonstrated. Experiments on the liquid–liquid extraction of Am(III) and Ln(III) nitrates were carried out to demonstrate advantageous features of chlorinated and fluorinated pyrrolidine diamides.

Graphical Abstract

1. Introduction

Lutetium is the heaviest and the smallest element of all the lanthanides [1,2]. Natural lutetium exists in two isotopic forms—in the form of a stable isotope 175Lu and a radioactive isotope 176Lu, which is used for radioisotope dating [3]. 177Lu is one of the most attractive isotopes in clinical nuclear medicine for the diagnosis and therapy of cancer [4].
A regular decrease in the ionic radii is observed in the lanthanide row. This effect, called “lanthanide contraction”, has been well studied and documented [5,6,7,8,9]. A variety of ligands have been shown to form complexes with lanthanides and the importance of lanthanide contraction on the structure of the formed complexes has been demonstrated [10,11,12]. Recently, we observed the effect of lanthanide contraction on tetrabutyl 1,10-phenanthroline-2,9-diamide [13], which acts as an N,N,O,O-tetradentate ligand. We revealed that in the La, Nd, and Eu complexes, metal ions are located below the plane of the phenanthroline core and the coordination number of the metal in the complexes with the ligand is 10. In the lutetium complex, the Lu3+ ion is located almost in the plane of the nucleus. As a result, the coordination number in the complex with lutetium decreases to nine and one of the nitrates became a monodentate. We have obtained a complex of lutetium nitrate with similar ligand-containing chlorine atoms at positions four and seven of the phenanthroline core. In this case, the coordination number of lutetium was also equal to nine. However, one of the nitrate groups was replaced with a more compact water molecule. The complex has been built on the principle of a tight ion pair having one nitrate group in the external coordination sphere [14].
Pyrrolidine-derived diamides are more compact ligands in comparison with tetrabutyl ligands. As a result, valuable properties of these ligands during the extraction separation of f-elements have been demonstrated [14,15,16,17]. Both the selectivity and extraction ability of the ligands under study decrease with the expansion of alicycle in the amide function despite an increase in solubility. These unusual properties can be explained by the molecular dynamics of such compounds [18,19,20,21,22].
It was shown in [17] that the presence of chlorine atoms in positions four and seven of the phenanthroline core and the variation of the structure of amide substituents may strongly affect the extraction properties of the ligands toward separation of lanthanides. At the same time, the replacement of chlorine atoms with fluorine allows us to expect increased resistance of ligands toward radiolysis in comparison with the chlorine-containing ligands. Recently, we reported on the effective synthesis of fluorinated phenanthrolinediamides [23]. A detailed comparison of the coordination properties of pyrrolidine-derived phenanthroline diamides is of significant interest. This study is devoted to a comparison of three phenanthroline-derived ligands L1L3 toward formation of lutetium complexes.

2. Results and Discussion

2.1. Structure of Ligands and Complexes

The ligands L1L3 have significant differences in their electronic properties (Figure 1). Calculated electrostatic potential maps (ESP maps) of ligands L1L3 permit the visualization of differences in electron distribution, which can be explained by the electron withdrawing influence of chlorine and fluorine.
Another possible way to view the electronic distribution in the ligands L1L3 is to use Merz-Kollman (ESP) charges [24]. For example, a difference in the corresponding charges on the nitrogen and oxygen atoms was observed for ligands L1L3 according to the calculated data (Table 1). The differences in the charges on nitrogen and oxygen atoms are due to the fact that the ligand structures are asymmetric. Ligands of this type exist in twisted conformations (Figure 1, a single potential scale from −0.02 to +0.02 conventional units). All quantum chemical calculations were performed by DFT with functional—B3LYP and basis—6-31G (d,p), using the Gaussian 16 program [25] method. The calculated ESP charges in carbonyl oxygen and phenanthroline nitrogen atoms in ligands L1 and L3 look very similar; small differences in the values of the corresponding charges are observed only for the ligand L2 (Table 1).
It is quite remarkable that the negative charges on the nitrogen atoms of the phenanthroline core are the highest in the L3 ligand, while the charges on the oxygen atoms in L3 are somewhat lower than in L1, but higher than in L2. As is well known, fluorine is a strong σ-electron acceptor, but at the same time a good π-donor. The effect of direct polar conjugation manifests itself most clearly on nitrogen atoms in para-positions with respect to fluorine atoms, which is reflected in the stability constants of complexes with lanthanide cations (see below).
As follows from the X-ray diffraction data (Figure 2), the carboxyl groups in all three L1L3 [15,17,23] ligands are on the same side of the phenanthroline core. The length of the C = O bonds varies from 1.220(3) Å in the ligand L2 to 1.246(2) Å in the ligand L1. Some geometric parameters are given in Table 2.
Next, we synthesized complex compounds of ligands L1L3 with lutetium nitrate (see the Materials and Methods section and Supplementary Materials Figures S1–S6). In each of the cases, we were able to investigate the structure of the obtained complex compounds by combining various spectral methods (NMR and IR) and by XRD analysis (Figure 3). Table 3 shows the IR data of ligands L1L3 and their complexes with Lu(NO3)3. It can be seen from the data that moving from L1 to L3 results in the shift difference of C=O increasing to 20 cm−1.
Single crystals were obtained by slow isothermal recrystallization of these complexes from the MeCN/CHCl3 solvent system. The resulting crystals of L2•Lu(NO3)3 and L3•Lu(NO3)3 also contained a lattice acetonitrile molecule. In the complexes with the ligands L1 and L2, the coordination number of the metal ion was nine (Figure 3).
In L1•Lu(NO3)3, the coordination environment of the metal ion includes the ligand L1 and two nitrate anions acting as bidentate ligands (Table 4) while the third nitrate group is displaced by the water molecule. The O-H…O hydrogen bond (O…O 2.683(5) Å, OHO 170.2(2)°) holds the latter species together to produce a centrosymmetric dimer from the neighboring complex molecules (Figure 3). No stacking interactions occur despite the presence of the extended aromatic phenanthroline core in the ligand L1; apart from the above hydrogen bonds, only weak van der Waals interactions operate in the crystal of this complex.
In L2•Lu(NO3)3, two nitrate anions occur in the outer sphere so that the metal ion coordinates one nitrate anion in a similar bidentate manner and three water molecules (Figure 3). The latter three molecules are involved in hydrogen bonding with the outer-sphere nitrate anions (O…O 2.72(2)–3.264(6) Å, OHO 126.3(3)–175.7(3)°) and the lattice acetonitrile molecule (O…N 2.853(6) Å, OHN 177.5(3)°) to produce corrugated double tapes along the crystallographic axis a (Figure 4). These tapes are packed into the 3D framework by Cl…Cl halogen bonds [26], as judged by the Cl…Cl distance of 3.251(2) Å and the C-Cl…Cl angle of 166.97(13)°, and stacking interactions between the parallel phenanthroline cores (the interplane angle is 0° and the intercentroid and shift distances are 4.6063(18) and 3.065(3) Å, respectively).
The coordination environment of the metal ion in L3•Lu(NO3)3 is formed only by the ligand L3 and three nitrate anions that all act as bidentate ligands. As a result, the coordination number reaches 10 (Figure 3). This type of coordination of the lutetium ion has already been observed in complexes with some organic ligands [27,28,29], however, it is still quite rare. For 1,10-phenanthroline-2,9-diamides, this is the first time that such a coordination number has been observed. As there are no convenient proton donors, such as the coordinated water molecules in L1•Lu(NO3)3 and L2•Lu(NO3)3, the crystal packing of L3•Lu(NO3)3 is dominated by parallel-displaced stacking interactions between the phenanthroline cores (the interplane angle is 0° and the intercentroid and shift distances are 4.217(5) and 2.507(8) Å, respectively) that produce centrosymmetric dimers (Figure 4).
In the complexes L1•Lu(NO3)3 and L3•Lu(NO3)3, the metal ion is located almost in the plane of the phenanthroline core of the ligand; the distance between the plane and the metal ion for these two complexes is 0.090(4) and 0.027(7) Å, respectively. In the complex L2•Lu(NO3)3, however, the metal ion moves away from the phenanthroline plane by 0.438(2) Å (Table 4). The Lu-N bond lengths increase in the row of substituents H–Cl–F in the four and seven positions of the ligands.

2.2. UV-Vis Titration

Stability constants for L3 with four representatives of lanthanides La(III), Nd(III), Eu(III), and Lu(III) were determined using UV-vis spectrophotometry titration. The absorption band of L3 in the range from 240 to 360 nm is very sensitive to its coordination environment.
Thus, these spectral changes can be used to determine stability constants (log β). Figure 5 shows an example of the absorption spectrum obtained by spectrophotometric titration of the ligand L3 with lanthanum trinitrate in a solution of dry acetonitrile (Figure 5a), the molar absorptions of free ligand L3 and Eu(III) complexes calculated from spectral deconvolution (Figure 5b), and the titration curve at maximum absorption (Figure 5c). Similar data for Nd(III), Eu(III), and Lu(III) are shown in Supplementary Materials Figure S7. All spectra have similar trends of change. With an increase in the number of metal ions added, the peak of the ligand L3 (~260 nm) gradually decreases, and a new peak corresponding to the metal-ligand complex appears in the 280 nm region. The obtained titration curves were analyzed using the Hypspec2014 program [30]. For all studied lanthanide ions, it was found that the titration data best correspond to the formation of L3:metal complexes 1:1 and 2:1 stoichiometry, which was previously also observed for L1 and L2 ligands [16,17]. The obtained stability constants of complexes of L3 with lanthanide ions (log β) are shown in Table 5. The L3 ligand has a greater affinity for light lanthanides since a decrease in log β1 is observed during the transition from La to Lu. It is worth noting that the stability constants for the L3 ligand with La3+, Nd3+, and Eu3+ are approximately 0.5 greater than the stability constants for the L1 and L2 ligands, while the log β1 values for complexes with Lu3+ are similar for all three ligands (log β1 = 5.98 ± 0.02 for L1 and 6.06 ± 0.02 for L2). The reason for this has been discussed above.

2.3. NMR Titration

Next, we studied the complexation of L3 with lutetium trinitrate at 25 °C using 1H and 19F NMR titration in deuterated acetonitrile (Figure 6). The gradual addition of a solution of lutetium pentahydrate Lu(H2O)5(NO3)3 in CD3CN to a solution of L3 in the same solvent led to a downfield shift of phenanthroline H5,6 and H3,8 signals as well as signals of the α-CH2 groups of the pyrrolidine moieties. Adding two equimoles of lutetium trinitrate caused the H5,6 and H3,8 signals to shift by 0.18 and 0.46 ppm, respectively, and the signals of the α-CH2 groups to shift by 0.17 and 0.22 ppm. An especially significant shift was observed in the 19F NMR. For example, the starting ligand had a signal with a shift of −112.22 ppm, which decreased to −99.25 ppm and −98.95 ppm after the addition of one or two equimoles of lutetium trinitrate, respectively. At the same time, the spectrum continued to change slightly even with the addition of lutetium nitrate to the solution of the 1:1 stoichiometry complex. In the case of stoichiometry, clear signals are observed in both the 1H and 19F NMR spectra when the metal:ligand ratio is 2:1, which may indicate that in the case of 1:1 stoichiometry, the ligand molecules may not be equivalent [31].
Having identified the structural differences in the studied lutetium complexes, the question of whether the ligands L2 and L3 are able to produce complexes with a metal:ligand ratio of 1:2 arose. We attempted to synthesize such a complex by reacting two ligand equivalents with lutetium nitrate. With the L2 ligand, we again obtained a complex with a metal:ligand raio of 1:1. With the ligand L3, a poly-nuclear oxo-complex (L3)3Lu3O2(NO3)5 in which the monomers are interconnected via the Lu-O bond, was obtained (Figure 7). In this complex, there are two symmetry-independent metal ions, the one in the center—Lu(2)—with a coordination number of eight and the two at the periphery—Lu(1)—with a coordination number of nine. The coordination environment of the former is formed by the ligand L3, with one nitrate anion coordinated in a bidentate manner and two oxygen atoms in the axial positions (Table 6). The latter coordinate involves another nitrate anion instead of one of the oxygen atoms. The central metal ion is located almost in the plane of the phenanthroline core of the ligand L3 (the distance from the plane and the metal ion is only 0.053(15) Å), while the metal ion at the periphery is shifted by 0.207(4) Å from this plane away from the central metal ion (Table 6).
In addition to the above coordinate bonds, the complex (L3)3Lu3O2(NO3)5 is stabilized by intramolecular parallel-displaced stacking interactions between the phenanthroline cores; the appropriate interplane angles are 5.27(14) and 6.0(2)°, the intercentroid distances are 4.476(4) and 3.662(5) Å, and the shift distances are 2.663(4) and 2.974(6) Å. Therefore, only weak van der Waals interactions pack the complex molecules into the 3D framework.

2.4. Extraction of Am(III) and Eu(III)

For ligands L2 and L3, the distribution ratio and separation factors of the Am(III)/Eu(III) pair were determined during extraction from nitric acid solutions with 0.01 M ligand solutions in 3-nitrobenzotrifluoride (F3) (Figure 8, Table 7). Ligand L1 did not participate in extraction experiments since this ligand has a negative log P = −0.08 and, as we have shown earlier, the distribution coefficients of Am(III) and Eu(III) during extraction with this ligand decrease with increasing phase contact time, which may be due to the fact that the complexes transition to the aqueous phase. The introduction of electron σ-acceptor fluorine atoms, which are stronger than chlorine atoms, leads to a greater decrease in the electron density on phenanthroline nitrogen atoms. As a result, the distribution ratios of Am(III) and Eu(III) are 5–10 times smaller in the case of extraction with L3 compared with L2.
However, the introduction of electron σ-acceptor atoms (both chlorine and fluorine) into the positions four and seven of the phenanthroline fragment leads to a decrease in the Brønsted basicity of diamides for ligands L2 and L3. As a result, an increase in the distribution coefficients of Am(III) and Eu(III) is observed over the entire studied range of nitric acid concentrations (1–6 mol·L−1). Unsubstituted diamides usually have an extraction maximum at a nitric acid concentration of 3–4 mol·L−1. A higher acid content leads to ligand protonation and a decrease in the distribution coefficients [17,32].
The extraction of lanthanides(III) (except for Pm(III)) was also studied. Figure 8b shows the distribution coefficients of the lanthanides(III) in the extraction of 0.01 M L2 and L3 from 5 M HNO3. The distribution coefficients of lanthanides(III) also take smaller values in the case of extraction with diamide with more σ-electron-withdrawing fluorine atoms, like what is seen with L3. Since the distribution coefficients of lanthanides(III) and americium(III) in the extraction of L3 under these conditions take values less than one, we also studied the extraction of lanthanides(III) with 0.05 M solutions of L3 in F3 from 3 and 5 M HNO3. During extraction from 5 M HNO3, the distribution coefficients of lanthanides(III) also turned out to be less than one, while the distribution coefficient of Am(III) was greater than one, which made it possible to separate it from lanthanides(III). It should be noted that in all cases, Am(III) was extracted better than lanthanides(III). The dependence of the distribution coefficients D on the cation atomic number Z when moving along the lanthanide series has a V-shaped character, which has already been observed for ligands of this type. The reasons for the appearance of such dependencies were considered by us earlier [14].

3. Materials and Methods

Chemical reagents such as Lu(NO3)3·xH2O and other inorganic/organic reagents and solvents were of analytical grade. Lutetium trinitrate hydrate Lu(NO3)3·xH2O was purchased from Sigma-Aldrich, Co. (St. Louis, MO, USA) and used without further purification. The water content x in lutetium nitrate was determined as x = 5. Deuterated solvent CD3CN for NMR spectra registration was purchased from Cambridge Isotope Laboratories, Inc. (Andover, MA, USA) and used without further purification. Acetonitrile, chloroform, and diethyl ether which were used for the synthesis of complexes were purified according to known procedures. C6F6 was purchased from Sigma-Aldrich, Co. (St. Louis, MO, USA) and used without further purification. Analytical grade 3-nitrobenzotrifluoride (F3) was purchased from Rhodia (France).
NMR spectra were recorded using standard 5 mm sample tubes on an Agilent 400-MR spectrometer (Agilent Technologies, Santa Clara, CA, USA) with operating frequencies of 400.1 MHz (1H) and 376 MHz (19F). IR spectra in the solid state were recorded on a Nicolet iS5 FTIR spectrometer (Thermo Fisher Scientific, Waltham, MA, USA) using an internal reflectance attachment with a diamond optical element and an attenuated total reflection (ATR) with a 45° angle of incidence. The resolution was 4 cm−1 and the number of scans was 32. HRMS ESI mass spectra were recorded on the MicroTof Bruker Daltonics and Orbitrap Elite instruments.
All quantum chemical calculations were performed by the Gaussian 16 program [25] DFT method with B3LYP functional and basis 6-31G (d,p).

3.1. Synthesis and Analytical Data

Syntheses of ligands L1, L2 and L3 were reported earlier [15,17,23].
General procedure for preparation of the complexes with Lu(NO3)3.
A solution of lutetium nitrate (0.1 mmol) in acetonitrile (1 mL) was added dropwise to a solution of 1,10-phenanthroline-2,9-dicarboxamide (0.1 mmol) in chloroform (1 mL) . After, the reaction mixture was concentrated in a vacuum to 1/10 of its initial volume and then treated with diethyl ether (2 mL). The resulting complex was filtered and washed with ether and dried in air.
N2,N9-bis(pyrrolidine)-N2,N9-diethyl-1,10-phenanthroline-2,9-dicarboxamide lutetium trinitrate L1•Lu(NO3)3. Yield 88.7% (65.2 mg). White powder. Tdecomp. > 340 °C; 1H NMR (CD3CN) δ 8.93 (d, J = 8.6 Hz, 2H, Phen), 8.59 (d, J = 8.6 Hz, 2H, Phen), 8.29 (s, 2H, Phen), 4.19 (t, J = 6.9 Hz, 4H, Pyrr), 3.89 (t, J = 6.9 Hz, 4H, Pyrr), 2.21–2.14 (m, 4H, Pyrr) 2.08–2.01 (m, 4H, Pyrr); IR (ν, cm−1) 3149, 3069, 2975, 2878 (C-H stretching vibrations), 1602 (C=O); HRMS (ESI-TOF) (m/z) [M + H+] calculated for [C22H22LuN6O8]+ 673.0902, found 673.0886.
N2,N9-bis(pyrrolidine)-4,7-dichloro-N2,N9-diethyl-1,10-phenanthroline-2,9-dicarboxamide lutetium trinitrate L2•Lu(NO3)3. Yield 88.1% (70.5 mg). Yellow powder. Tdecomp. 240 °C; 1H NMR (CD3CN) δ 8.62 (s, 2H, Phen), 8.58 (s, 2H, Phen), 4.18 (t, J = 6.9 Hz, 4H, Pyrr), 3.88 (t, J = 6.9 Hz, 4H, Pyrr), 2.20–2.13 (m, 4H, Pyrr), 2.07–2.00 (m, 4H, Pyrr); IR (ν, cm−1) 3120, 3083, 2987, 2883 (C-H stretching vibrations), 1609 (C=O); HRMS (ESI-TOF) (m/z) [M + H+] calculated for [C22H20Cl2LuN6O8]+ 741.0122, found 741.0122
N2,N9-bis(pyrrolidine)-4,7-difluoro-N2,N9-diethyl-1,10-phenanthroline-2,9-dicarboxamide lutetium trinitrate L3•Lu(NO3)3. Yield 96.8% (74.6 mg). White powder. Tdecomp. 250 °C; 1H NMR (CD3CN) δ 8.40 (s, 2H, Phen), 8.33 (d, J = 9.8 Hz, 2H, Phen), 4.16 (t, J = 6.9 Hz, 4H, Pyrr), 3.87 (t, J = 6.9 Hz, 4H, Pyrr), 2.21–2.12 (m, 4H, Pyrr), 2.07–1.99 (m, 4H, Pyrr); 19F NMR (376 MHz, Acetonitrile-d3) δ −99.52 (d, J = 9.8 Hz), −103.28; IR (ν, cm−1) 3115, 3087, 2979, 2881 (C-H stretching vibrations), 1608 (C=O); HRMS (ESI-TOF) (m/z) [M + H+] caclulated for [C22H20F2LuN6O8]+ 709.0713, found 709.0705
Oxo-complex (L3)3Lu3O2(NO3)5 was obtained in accordance with the general procedure starting from 0.2 mmol of L3 and 0.1 mmol of lutetium nitrate. The treatment of the residue with ether yielded 112.8 mg of yellowish powder. This material was used for growing single crystals.

3.2. UV-Vis Titration Experiment

UV−VIS spectra were recorded at a temperature of 25.0 ± 0.1 °C in the wavelength region of 200−500 nm (0.5 nm interval) on a Shimadzu UV 1800 spectrophotometer controlled by LabSolutions version 1.2.35 UV-Vis software with a thermostatic attachment (Shimadzu TCC-100) using quartz cuvettes with an optical path length of 10 mm. A stock solution of the ligand was prepared (ca. 10−4 mol L−1) by dissolving the respective ligand in CH3CN; a working ligand solution (ca. 10−5 mol L−1) was then prepared from the initial solution. A working titrant solution (10−3 mol L−1) was prepared by dissolving the respective lanthanide(III) nitrate hydrate Ln(NO3)3·xH2O compound (Ln = La, Nd, Eu, Lu; x = 6 in the case of La, Nd, Eu and x = 5 in the case of Lu) in CH3CN. Acetonitrile (CH3CN; 99.95%, HPLC grade, Panreac AppliChem) was dried over molecular sieves (zeolite KA, 3 Å, balls, diameter 1.6−2.5 mm, production HKC Corp., Hong Kong) prior to use. The titration was carried out by adding 2 μL aliquots of the working metal cation solution to 2 mL of the working ligand solution in the titration cell. The titration continued until no obvious change was observed in the spectra. The stability constants of the Ln(III) complexes were calculated using the HypSpec2014 program.

3.3. NMR Titration Experiment

For the NMR titration experiments, the stock solutions of L3 and Lu(NO3)3 were obtained by dissolving weighed amounts of ligand L3 and Lu(NO3)3·5H2O in CD3CN, respectively. The initial aliquot solution of L3 was divided equally into five NMR tubes equally and a certain amount of Lu(NO3)3 solution was added to four of them to obtain a series of samples with molar ligand:metal ratios of 1:0, 1:0.5, 1:1, 1:1.5, and 1:2. The total concentrations of metals and ligands were 0.0585 mol L−1 and 0.0244 mol L−1 , respectively. After the addition, 1H and 19F NMR spectra were recorded on an Agilent 400-MR instrument with a frequency of 400.1 MHz (1H) and 376 MHz (19F). C6F6 in CD3CN was used as a reference for the 19F-NMR spectra.

3.4. X-ray Crystallography

X-ray diffraction data for the ligand L3 and for the complexes L•Lu(NO3)3 were collected at 295 K using a STOE STADIVARI diffractometer with a Pilatus 100K detector using graphite monochromated Mo-Kα radiation (λ = 0.71073 Å) for L2•Lu(NO3)3 and the Cu Kα radiation (λ = 1.54186 Å) from the fine-focus Cu GeniX 3D radiation source for the others. The data for (L3)3Lu3O2(NO3)5 were collected at 100 K with a Bruker Quest D8 CMOS diffractometer using graphite monochromated Mo-Kα radiation (λ = 0.71073 Å). The structures of the ligand L3 and of the complexes L•Lu(NO3)3 were solved and refined with the program SHELX [33], while the ShelXT [34] structure solution program using intrinsic phasing and refined with the XL [33] refinement package using least-squares minimization against F2 was used for (L3)3Lu3O2(NO3)5. The non-hydrogen atoms were refined by using the anisotropic full matrix least-squares procedure. Hydrogen atoms of coordinated water molecules were located from different Fourier syntheses while the positions of others were calculated, and all of them were refined in isotropic approximation within the riding model. Crystal data and structure refinement parameters for the ligand L3 and the complexes L•Lu(NO3)3 and (L3)3Lu3O2(NO3)5 are given in Table S1 (see the Supplementary Materials).
CCDC 2221666 (for L3), 2221667 (for L1•Lu(NO3)3), 2191901 (for L2•Lu(NO3)3), 2221670 (for L3•Lu(NO3)3), and 2232270 (for (L3)3Lu3O2(NO3)5) contain the supplementary crystallographic data for this paper.

4. Conclusions

In summary, we revealed important structural features of complexes of lutetium trinitrate with pyrrolidine-derived phenanthroline diamides L1L3. Depending on the presence and type of halide atoms in positions four and seven of the phenanthroline core, the coordination number of lutetium can be equal to either nine (complexes of L1 and L2), which is typical for lutetium complex compounds, or 10 (fluorine-containing ligand L3), which is very rare. The measurement of bond lengths in the structures of complexes studied by the XRD analysis shows that in the case of L3, the bonds of the lutetium atom with the coordination centers are longer than similar bonds in the complexes of the other two ligands. Also, in the case of the lutetium trinitrate complex with L3, the smallest out-of-plane distance of the metal from the phenanthroline plane is observed. Apparently, this makes the coordination of three bidentate nitrate groups possible. It is also noteworthy that in the case of a complex with the L2 ligand, only one of three nitrate groups is located in the internal coordination sphere, while the other two nitrate groups are displaced by three compact water molecules.
Using UV-Vis and NMR titration techniques, we elucidated the coordination chemistry of ligands in acetonitrile solutions and obtained the values of stability constants. The log β2 values for lanthanide complexes of L3 turned out to be much higher than those of the corresponding complexes of ligands L1 and L2. Counting on greater stability of the L3:lanthanide = 2:1 complexes, we tried to obtain complexes of such stoichiometry with lutetium trinitrate in an individual form. As a result, the oxo-complex was isolated using XRD analysis. It is very likely that in this case, one ligand molecule works as a base by binding the nitrate anion.
In addition, we performed preliminary liquid–liquid extraction tests. The results obtained indicate a lower extraction efficiency of the ligand L3 compared to its chlorine-containing analog L2, but the selectivity in the separation of the Am(III)/Eu(III) remains high. This is probably the result of the lower lipophilicity of the L3 ligand, which affects the formation of water-soluble complexes with lanthanide nitrates.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/ijms24065569/s1.

Author Contributions

Conceptualization, N.A.A. and P.S.L.; investigation, N.A.A., M.V.E. and P.I.M.; synthesis, N.A.A.; XRD analysis, V.A.T., S.A.A. and Y.V.N.; data curation, P.S.L. and V.G.P.; writing—original draft preparation, N.A.A., M.V.E., P.I.M. and Y.V.N.; writing—review and editing, P.S.L., A.V.Y., Y.V.N. and X.B.; supervision, Y.A.U. and V.G.N.; funding acquisition, P.S.L. and Y.A.U. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by Russian Science Foundation (grant no. 21-73-10067).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Samples of the compounds are not available from the authors.

Acknowledgments

The authors acknowledge support from the M. V. Lomonosov Moscow State University Program of Development. We appreciate B.N. Tarasevich for IR-spectra registering and M.F. Vokuev for registering HRMS. We appreciate V.A. Chertkov for providing the Gaussian 16 program. The X-ray diffraction data for the complex (L3)3Lu3O2(NO3)5 were collected using the equipment of the Centre for Molecular Composition Studies of INEOS RAS supported by the Ministry of Science and Higher Education of the Russian Federation (Contract/agreement No. 075-00697-22-00).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Greenwood, N.N.; Earnshaw, A. Chemistry of the Elements, 2nd ed.; Elsevier Butterworth-Heinemann: Oxford, UK, 1997. [Google Scholar]
  2. Cotton, S. Lanthanide and Actinide Chemistry; John Wiley & Sons, Ltd.: Colchester, UK, 2006. [Google Scholar]
  3. Mignerey, A.C. Dating Techniques; Department of Chemistry and Biochemistry, University of Maryland: College Park, MA, USA, 2003. [Google Scholar] [CrossRef]
  4. Kurenkov, N.V.; Shubin, Y.N. Radionuclides for nuclear medicine. Med. Radiol. Radiat. Saf. 1996, 41, 54–63. [Google Scholar]
  5. Cotton, S.A.; Raithby, P.R. Systematics and surprises in lanthanide coordination chemistry; Coordination. Chem. Rev. 2017, 340, 220–231. [Google Scholar] [CrossRef]
  6. Quadrelli, E.A. Lanthanide Contraction over the 4f Series Follows a Quadratic Decay. Inorg. Chem. 2002, 41, 167–169. [Google Scholar] [CrossRef] [PubMed]
  7. Raymond, K.N.; Wellman, D.L.; Sgarlata, C.; Hill, A.P. Curvature of the lanthanide contraction: An explanation. Comptes Rendus Chim. 2010, 13, 849–852. [Google Scholar] [CrossRef] [Green Version]
  8. Lundberg, D.; Persson, I. The size of actinoid(III) ions—Structural analysis vs. common misinterpretations. Coord. Chem. Rev. 2016, 318, 131–134. [Google Scholar] [CrossRef] [Green Version]
  9. Seitz, M.; Oliver, A.G.; Raymond, K.N. The Lanthanide Contraction Revisited. J. Am. Chem. Soc. 2007, 129, 11153–11160. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  10. Yue, Z.; Lu, H.; Li, Z.; Guo, S.; Song, J.; Ren, Y.M.; Huang, Y.; Lin, J.; Wang, J. Structural Evolution and Tuneable Photoluminescence of f- Element Bearing Coordination Polymers of the 2,4,6-tri-α-pyridyl-1,3,5-triazine Ligand. CrystEngComm 2019, 21, 5059–5066. [Google Scholar] [CrossRef]
  11. Drew, M.G.B.; Hudson, M.J.; Iveson, P.B.; Madic, C.; Russell, M.L. A study of lanthanide complexes formed with the terdentate nitrogen ligand 4-amino-bis(2,6-(2-pyridyl))-1,3,5-triazine. Relevance to the separation of actinides and lanthanides by solvent extraction. J. Chem. Soc. Dalton Trans. 2000, 16, 2711–2720. [Google Scholar] [CrossRef]
  12. Cotton, S.A.; Franckevicius, V.; Mahon, M.F.; Ooi, L.L.; Raithby, P.R.; Teat, S.J. Structures of 2,4,6-tri-a-pyridyl-1,3,5-triazine complexes of the lanthanoid nitrates: A study in the lanthanoid contraction. Polyhedron 2006, 25, 1057–1068. [Google Scholar] [CrossRef]
  13. Lemport, P.S.; Evsiunina, M.V.; Nelyubina, Y.V.; Isakovskaya, K.L.; Khrustalev, V.N.; Petrov, V.S.; Pozdeev, A.S.; Matveev, P.I.; Ustynyuk, Y.A.; Nenajdenko, V.G. Significant impact of lanthanide contraction on the structure of the phenanthroline complexes. Mendeleev Commun. 2021, 31, 853–855. [Google Scholar] [CrossRef]
  14. Ustynyuk, Y.A.; Zhokhova, N.I.; Gloriozov, I.P.; Matveev, P.I.; Evsiunina, M.V.; Lemport, P.S.; Pozdeev, A.S.; Petrov, V.G.; Yatsenko, A.V.; Tafeenko, V.A.; et al. Competing Routes in the Extraction of Lanthanide Nitrates by 1,10-Phenanthroline-2,9-diamides: An Impact of Structure of Complexes on the Extraction. Int. J. Mol. Sci. 2022, 23, 15538. [Google Scholar] [CrossRef] [PubMed]
  15. Krapcho, A.P.; Ali, A. Synthesis of 2,9-diacyl-1, 10-phenanthrolines. J. Heterocycl. Chem. 2004, 41, 795–798. [Google Scholar] [CrossRef]
  16. Lemport, P.S.; Evsunina, M.V.; Matveev, P.I.; Petrov, V.S.; Pozdeev, A.S.; Khult, E.K.; Nelyubina, Y.V.; Isakovskaya, K.L.; Roznyatovsky, V.A.; Gloriozov, I.P.; et al. 2-Methylpyrrolidine Derived 1,10-Phenanthroline-2,9-diamides: Promising Extractants for Am(III)/Ln(III) Separation. Inorg. Chem. Front. 2022, 9, 4402–4412. [Google Scholar] [CrossRef]
  17. Lemport, P.S.; Matveev, P.I.; Yatsenko, A.V.; Evsiunina, M.V.; Petrov, V.S.; Tarasevich, B.N.; Roznyatovsky, V.A.; Dorovatovskii, P.V.; Khrustalev, V.N.; Zhokhov, S.S.; et al. The impact of alicyclic substituents on the extraction ability of new family of 1,10-phenanthroline-2,9-diamides. RSC Adv. 2020, 10, 26022–26033. [Google Scholar] [CrossRef] [PubMed]
  18. Borisova, N.E.; Reshetova, M.D. Quantum chemical modeling of 2,2´-bipyridine-6,6´-dicarboxylic acid diamide structures: A relationship between the extraction ability and conformational behavior of the ligands. Russ. Chem. Bull. 2015, 64, 1882–1890. [Google Scholar] [CrossRef]
  19. Lewis, F.W.; Harwood, L.M.; Hudson, M.J.; Drew, M.G.B.; Hubscher-Bruder, V.; Videva, V.; Arnaud-Neu, F.; Stamberg, K.; Vyas, S. BTBPs versus BTPhens: Some Reasons for Their Differences in Properties Concerning the Partitioning of Minor Actinides and the Advantages of BTPhens. Inorg. Chem. 2013, 52, 4993–5005. [Google Scholar] [CrossRef]
  20. Xu, L.; Pu, N.; Li, Y.; Wei, P.; Sun, T.; Xiao, C.; Chen, J.; Xu, C. Selective Separation and Complexation of Trivalent Actinide and Lanthanide by a Tetradentate Soft–Hard Donor Ligand: Solvent Extraction, Spectroscopy, and DFT Calculations. Inorg. Chem. 2019, 58, 4420–4430. [Google Scholar] [CrossRef]
  21. Jansone-Popova, S.; Ivanov, A.S.; Bryantsev, V.S.; Sloop, F.V.; Custelcean, R.; Popovs, I.; Dekarske, M.M.; Moyer, B.A. Bis-lactam-1,10-phenanthroline (BLPhen), a New Type of Preorganized Mixed N,O-Donor Ligand That Separates Am(III) over Eu(III) with Exceptionally High Efficiency. Inorg. Chem. 2017, 56, 5911–5917. [Google Scholar] [CrossRef] [PubMed]
  22. Lumetta, G.J.; Rapko, B.M.; Garza, P.A.; Hay, B.P.; Gilbertson, R.D.; Weakley, T.J.R.; Hutchison, J.E. Deliberate Design of Ligand Architecture Yields Dramatic Enhancement of Metal Ion Affinity. J. Am. Chem. Soc. 2002, 124, 5644–5645. [Google Scholar] [CrossRef] [PubMed]
  23. Avagyan, N.A.; Lemport, P.S.; Lyssenko, K.A.; Gudovannyy, A.O.; Roznyatovsky, V.A.; Petrov, V.S.; Vokuev, M.F.; Ustynyuk, Y.A.; Nenajdenko, V.G. First Example of Fluorinated Phenanthroline Diamides: Synthesis, Structural Study, and Complexation with Lanthanoids. Molecules 2022, 27, 4705. [Google Scholar] [CrossRef] [PubMed]
  24. Besler, B.H.; Merz, K.M., Jr.; Kollman, P.A. Atomic charges derived from semiempirical methods. J. Comp. Chem. 1990, 11, 431–439. [Google Scholar] [CrossRef]
  25. Frisch, M.J.; Trucks, G.W.; Schlegel, H.B.; Scuseria, G.E.; Robb, M.A.; Cheeseman, J.R.; Scalmani, G.; Barone, V.; Petersson, G.A.; Nakatsuji, H.; et al. Gaussian 16, Revision C.01; Gaussian, Inc.: Wallingford, CT, USA, 2016. [Google Scholar]
  26. Erdélyi, M.; Esterhuysen, C.; Zhu, W. Halogen Bonding: From Fundamentals to Applications. ChemPlusChem 2021, 86, 1229–1230. [Google Scholar] [CrossRef] [PubMed]
  27. Savić, A.; Kaczmarek, A.M.; Deun, R.V.; Heck, K.V. DNA intercalating near-infrared luminescent lanthanide complexes containing dipyrido[3,2-a:2′,3′-c]phenazine (dppz) ligands : Synthesis, crystal structures, stability, luminescence properties and CT-DNA interaction. Molecules 2020, 25, 5309. [Google Scholar] [CrossRef] [PubMed]
  28. Kapert, D.L.; Semenova, L.I.; Sobolev, A.N.; White, A.H. Structural Systematics of Rare Earth Complexes. Tris(nitrato- O,O′)(bidentate- N, N′)lutetium(III), N, N′-Bidentate = 2,2′-Bipyridine or 1,10-Phenanthroline. Aust. J. Chem. 1996, 49, 1005–1008. [Google Scholar] [CrossRef]
  29. Drew, M.G.B.; Foreman, M.R.S.; Hudson, M.J.; Kennedy, K.F. Structural studies of lanthanide complexes with tetradentate nitrogen ligands. Inorg. Chim. Acta 2004, 357, 4102–4112. [Google Scholar] [CrossRef]
  30. Gans, P.; Sabatini, A.; Vacca, A. Investigation of equilibria in solution. Determination of equilibrium constants with the HYPERQUAD suite of programs. Talanta 1996, 43, 1739–1753. [Google Scholar] [CrossRef] [PubMed]
  31. Meng, R.; Xu, L.; Yang, X.; Sun, M.; Xu, C.; Borisova, N.E.; Zhang, X.; Lei, L.; Xiao, C. Influence of a N-Heterocyclic Core on the Binding Capability of N,O-Hybrid Diamide Ligands toward Trivalent Lanthanides and Actinides. Inorg. Chem. 2021, 60, 8754–8764. [Google Scholar] [CrossRef] [PubMed]
  32. Alyapyshev, M.; Ashina, J.; Dar’in, D.; Kenf, E.; Kirsanov, D.; Tkachenko, L.; Legin, A.; Starova, G.; Babain, V. 1,10-Phenanthroline-2,9-dicarboxamides as ligands for separation and sensing of hazardous metals. RSC Adv. 2016, 6, 68642. [Google Scholar] [CrossRef]
  33. Sheldrick, G.M. A short history of SHELX. Acta Crystallogr. 2008, A64, 112–122. [Google Scholar] [CrossRef] [Green Version]
  34. Sheldrick, G.M. SHELXT-Integrated space-group and crystal-structure determination. Acta Crystallogr. 2015, A71, 3–8. [Google Scholar] [CrossRef] [Green Version]
Figure 1. ESP maps of ligands L1L3 in two projections.
Figure 1. ESP maps of ligands L1L3 in two projections.
Ijms 24 05569 g001aIjms 24 05569 g001b
Figure 2. General view of ligands L1 (a), L2 (b), and L3 (c) in two projections. The non-hydrogen atoms are shown as thermal ellipsoids at a 50% probability level [16,17].
Figure 2. General view of ligands L1 (a), L2 (b), and L3 (c) in two projections. The non-hydrogen atoms are shown as thermal ellipsoids at a 50% probability level [16,17].
Ijms 24 05569 g002aIjms 24 05569 g002b
Figure 3. General view of the complexes L1•Lu(NO3)3 (a), L2•Lu(NO3)3 (b), and L3•Lu(NO3)3 (c) in two projections. Hydrogen atoms, except those of coordinated water molecules as well as the lattice solvent in L3•Lu(NO3)3, are omitted for clarity, and the non-hydrogen atoms are shown as thermal ellipsoids at a 50% probability level.
Figure 3. General view of the complexes L1•Lu(NO3)3 (a), L2•Lu(NO3)3 (b), and L3•Lu(NO3)3 (c) in two projections. Hydrogen atoms, except those of coordinated water molecules as well as the lattice solvent in L3•Lu(NO3)3, are omitted for clarity, and the non-hydrogen atoms are shown as thermal ellipsoids at a 50% probability level.
Ijms 24 05569 g003aIjms 24 05569 g003b
Figure 4. Fragments of the crystal packing illustrating the formation of a hydrogen-bonded dimer in L1•Lu(NO3)3 (a), a hydrogen-bonded tape in L2•Lu(NO3)3 (b), and a dimer formed by stacking interactions in L3•Lu(NO3)3 (c). Red dotted lines represent hydrogen bonds and the phenanthroline cores involved in stacking interactions are highlighted in pink.
Figure 4. Fragments of the crystal packing illustrating the formation of a hydrogen-bonded dimer in L1•Lu(NO3)3 (a), a hydrogen-bonded tape in L2•Lu(NO3)3 (b), and a dimer formed by stacking interactions in L3•Lu(NO3)3 (c). Red dotted lines represent hydrogen bonds and the phenanthroline cores involved in stacking interactions are highlighted in pink.
Ijms 24 05569 g004
Figure 5. Spectrophotometric titration of L3 (ca. 10−5 mol L−1) with La3+ ions (ca. 5 × 10−4 mol L−1) in CH3CN solution (T = 25.0 ± 0.1 °C, I = 0 M, V0 = 2.0 mL): (a) absorption spectra, (b) the molar absorptivities of free ligand L3 and La(III) complexes calculated from spectral deconvolution, and (c) titration curve at maximum absorption (282 nm).
Figure 5. Spectrophotometric titration of L3 (ca. 10−5 mol L−1) with La3+ ions (ca. 5 × 10−4 mol L−1) in CH3CN solution (T = 25.0 ± 0.1 °C, I = 0 M, V0 = 2.0 mL): (a) absorption spectra, (b) the molar absorptivities of free ligand L3 and La(III) complexes calculated from spectral deconvolution, and (c) titration curve at maximum absorption (282 nm).
Ijms 24 05569 g005
Figure 6. 19F (left) and 1H (right) spectra of NMR titration in CD3CN at 25 °C. From top to bottom (a) before adding Lu3+, (b) L3: Lu3+ = 1:0.5, (c) L3: Lu3+ = 1:1, (d) L3: Lu3+ = 1:1.5, and (e) L3: Lu3+ = 1:2.
Figure 6. 19F (left) and 1H (right) spectra of NMR titration in CD3CN at 25 °C. From top to bottom (a) before adding Lu3+, (b) L3: Lu3+ = 1:0.5, (c) L3: Lu3+ = 1:1, (d) L3: Lu3+ = 1:1.5, and (e) L3: Lu3+ = 1:2.
Ijms 24 05569 g006
Figure 7. General view of the oxo-complex (L3)3Lu3O2(NO3)5 with hydrogen atoms omitted for clarity and other atoms shown as thermal ellipsoids at a 50% probability level. Phenanthroline cores involved in stacking interaction are highlighted by pink.
Figure 7. General view of the oxo-complex (L3)3Lu3O2(NO3)5 with hydrogen atoms omitted for clarity and other atoms shown as thermal ellipsoids at a 50% probability level. Phenanthroline cores involved in stacking interaction are highlighted by pink.
Ijms 24 05569 g007
Figure 8. The dependence of the distribution ratio Am(III) and Eu(III) on the concentration of nitric acid in the equilibrium aqueous phase during extraction with 0.01 M ligand solutions in F3 (a). Distribution ratios of lanthanides(III) and Am(III) for extraction by (b) 0.01 mol L−1 L2 and L3 from 5 mol L−1 HNO3 and (c) 0.05 mol L−1 L3 from 3 and 5 mol L−1 HNO3.
Figure 8. The dependence of the distribution ratio Am(III) and Eu(III) on the concentration of nitric acid in the equilibrium aqueous phase during extraction with 0.01 M ligand solutions in F3 (a). Distribution ratios of lanthanides(III) and Am(III) for extraction by (b) 0.01 mol L−1 L2 and L3 from 5 mol L−1 HNO3 and (c) 0.05 mol L−1 L3 from 3 and 5 mol L−1 HNO3.
Ijms 24 05569 g008
Table 1. ESP charges of NPhen and Oamide atoms of ligands L1L3.
Table 1. ESP charges of NPhen and Oamide atoms of ligands L1L3.
LigandESP Charges
NPhenOamide
L1−0.339/−0.367−0.500/−0.502
L2−0.263/−0.307−0.488/−0.489
L3−0.366/−0.405−0.501/−0.502
Table 2. Some geometric parameters of ligands L1L3 [16,17].
Table 2. Some geometric parameters of ligands L1L3 [16,17].
Ligand/Bond Length, ÅC = O(1)C = O(2)C-N(Amide)C-N(Amide)OCCN(°)OCCN(°)
L11.241(2)
[1.246(2)]
1.241(2)
[1.245(2)]
1.347(2)
[1.337(2)]
1.348(2)
[1.344(2)]
139.48(16)
[150.64(18)]
160.72(16)
[173.73(18)]
L21.220(3)1.223(3)1.328(3)1.332(3)128.6(3)141.1(3)
L31.232(4)1.235(4)1.330(4)1.324(4)124.6(3)152.6(3)
The values for the second symmetry-independent molecule are given in brackets.
Table 3. IR data of ligands L1L3 and their complexes with Lu(NO3)3.
Table 3. IR data of ligands L1L3 and their complexes with Lu(NO3)3.
IR (CO), cm−1 IR (CO), cm−1Δ (CO), cm−1
L11613L1•Lu(NO3)3160211
L21622L2•Lu(NO3)3160913
L31628L3•Lu(NO3)3160820
Table 4. Some geometric parameters of the complexes L•Lu(NO3)3.
Table 4. Some geometric parameters of the complexes L•Lu(NO3)3.
L1•Lu(NO3)3L2•Lu(NO3)3L3•Lu(NO3)3
Bond Length, Å
RM-O(1)2.304(3)2.311(2)2.324(5)
RM-O(2)2.267(3)2.321(3)2.313(5)
RM-N(1)2.416(3)2.472(3)2.500(6)
RM-N(2)2.404(3)2.483(3)2.531(6)
RM-ONO2(2)2.363(3)–2.471(4)2.405(3), 2.478(3)2.406(7)–2.490(6)
RM-OH22.281(3)2.260(2)–2.392(3)-
Out-of-plane0.090(4)0.438(2)0.027(7)
Torsions (°)
Nphen-C-C=O1.8(6), 11.5(6)4.6(4), 6.0(4)24.4(11), 17.4(12)
Camide1-Namide-C=O a)173.2(5), 177.3(5)173.6(3), 167.3(5) [173.4(12)]172.0(8), 174.0(10)
Camide2-Namide-C=O a)0.6(7), 2.7(5)1.8(5), 6.2(6) [16.8(8)]8.5(13), 1.4(14)
a) The values for the minor component of the disordered moiety are given in brackets.
Table 5. Stability constants (log β) for L1, L2, and L3 complexes with Ln(III) nitrates.
Table 5. Stability constants (log β) for L1, L2, and L3 complexes with Ln(III) nitrates.
LigandMetal Ionlog β1log β2
L1La3+5.90 ± 0.0211.64 ± 0.04
Nd3+5.96 ± 0.0211.78 ± 0.04
Eu3+5.92 ± 0.0211.62 ± 0.04
Lu3+5.98 ± 0.0311.78 ± 0.04
L2La3+5.82 ± 0.0211.63 ± 0.05
Nd3+5.85 ± 0.0211.64 ± 0.05
Eu3+5.90 ± 0.0211.66 ± 0.05
Lu3+6.06 ± 0.0211.63 ± 0.04
L3La3+6.53 ± 0.0211.81 ± 0.06
Nd3+6.50 ± 0.0111.70 ± 0.03
Eu3+6.39 ± 0.0211.62 ± 0.04
Lu3+6.02 ± 0.0212.04 ± 0.04
Table 6. Some geometric parameters of the oxo-complex (L3)3Lu3O2(NO3)5 a).
Table 6. Some geometric parameters of the oxo-complex (L3)3Lu3O2(NO3)5 a).
Lu(1)Lu(2)
Bond length, Å
RM-O(1), RM-O(2)2.311(4), 2.320(4)2.34(2) [2.24(2)]
RM-N(1), RM-N(2)2.463(4), 2.420(4)2.431(7) [2.400(11)]
RM-ONO2(2)2.390(18)–2.474(4)2.375(3)
RM-O2.129(2)2.116(2)
Out-of-plane shift0.207(4)0.053(15)
Bond angles (°)
Lu-O-Lu170.99(14)
O-Lu-O-154.79(14)
Torsions (o)
Nphen-C-C=O12.1(6), 13.3(6)9.6(18) [10(2)]
Camide-Namide-C=O2.7(7)–3(4)1(2) [2(2)]
a) The values for the minor component of the disordered moiety are given in brackets.
Table 7. The values of the Am/Eu pair separation factors during extraction with 0.01 M ligand solutions in F3 from nitric acid solutions.
Table 7. The values of the Am/Eu pair separation factors during extraction with 0.01 M ligand solutions in F3 from nitric acid solutions.
C(HNO3), ML2L3
11.43.3
25.98.7
312.114.8
412.712.3
516.714.1
618.013.9
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

Avagyan, N.A.; Lemport, P.S.; Evsiunina, M.V.; Matveev, P.I.; Aksenova, S.A.; Nelyubina, Y.V.; Yatsenko, A.V.; Tafeenko, V.A.; Petrov, V.G.; Ustynyuk, Y.A.; et al. Pyrrolidine-Derived Phenanthroline Diamides: An Influence of Fluorine Atoms on the Coordination of Lu(III) and Some Other f-Elements and Their Solvent Extraction. Int. J. Mol. Sci. 2023, 24, 5569. https://doi.org/10.3390/ijms24065569

AMA Style

Avagyan NA, Lemport PS, Evsiunina MV, Matveev PI, Aksenova SA, Nelyubina YV, Yatsenko AV, Tafeenko VA, Petrov VG, Ustynyuk YA, et al. Pyrrolidine-Derived Phenanthroline Diamides: An Influence of Fluorine Atoms on the Coordination of Lu(III) and Some Other f-Elements and Their Solvent Extraction. International Journal of Molecular Sciences. 2023; 24(6):5569. https://doi.org/10.3390/ijms24065569

Chicago/Turabian Style

Avagyan, Nane A., Pavel S. Lemport, Mariia V. Evsiunina, Petr I. Matveev, Svetlana A. Aksenova, Yulia V. Nelyubina, Alexandr V. Yatsenko, Viktor A. Tafeenko, Vladimir G. Petrov, Yuri A. Ustynyuk, and et al. 2023. "Pyrrolidine-Derived Phenanthroline Diamides: An Influence of Fluorine Atoms on the Coordination of Lu(III) and Some Other f-Elements and Their Solvent Extraction" International Journal of Molecular Sciences 24, no. 6: 5569. https://doi.org/10.3390/ijms24065569

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