Next Article in Journal
Photoelectron Circular Dichroism as a Probe of Chiral Hydrocarbons
Next Article in Special Issue
Enantiopure Cyclometalated Rh(III) and Ir(III) Complexes Displaying Rigid Configuration at Metal Center: Design, Structures, Chiroptical Properties and Role of the Iodide Ligand
Previous Article in Journal
Nanoencapsulated Myricetin to Improve Antioxidant Activity and Bioavailability: A Study on Zebrafish Embryos
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Complexation Behavior of Pinene–Bipyridine Ligands towards Lanthanides: The Influence of the Carboxylic Arm

1
Haute Ecole d’Ingénierie et d’Architecture Fribourg, HEIA-FR, HES-SO, University of Applied Sciences of Western Switzerland, Pérolles 80, CH-1705 Fribourg, Switzerland
2
Department of Chemistry, University of Fribourg, Chemin du Musée 9, CH-1700 Fribourg, Switzerland
3
School of Chemistry & Chemical Engineering, Henan University of Technology, Zhengzhou 450001, China
*
Author to whom correspondence should be addressed.
Chemistry 2022, 4(1), 18-30; https://doi.org/10.3390/chemistry4010002
Submission received: 8 December 2021 / Revised: 21 December 2021 / Accepted: 27 December 2021 / Published: 31 December 2021

Abstract

:
The complexation behavior of two novel, chiral pinene–bipyridine-type ligands ((–)-HL1 and (–)-HL2) containing a carboxylic arm towards lanthanide Ln(III) (Ln = La, Eu, Lu) ions was investigated through spectroscopic methods. The association constants of the mononuclear complexes determined from the UV-Vis titrations indicated that the ligand (–)-HL1 possessing a shorter carboxylic arm formed more stable complexes compared with (–)-HL2, whose carboxylic arm had one more methylene unit. This is due to the formation of more stable seven-member metal chelate rings in the first case as compared with the eight-member metal chelate rings in the second. IR and fluorescence spectroscopy provided additional information about the structure of these complexes.

1. Introduction

The use of coordination chemistry for the development of large, discrete, supramolecular architectures by self-assembly is well known. The strength of the interaction and the directionality induced by the coordination bonds can be more effective in self-assembly processes than just H-bonding or van der Waals forces [1]. Indeed, a plethora of discrete, supramolecular architectures based on coordination bonds (e.g., coordination cages [2,3,4,5], or helicates [6]) have been obtained.
Often, the complexity of the self-assembled species increases with the coordination number of the metal ions. Such a behavior can be expected for the lanthanide ions Ln(III) due to their large ionic radii and high coordination numbers. Moreover, their interesting features such as large Stokes shifts, emission [7], and magnetic properties [8] can lead to the formation of self-assembled architectures with interesting functionalities [9]. Such Ln(III) complexes have found many applications in lanthanide–actinide extraction processes [10,11], bioimaging [12], sensing [13], lighting and displays [14,15,16,17,18], or as single-molecular magnets [8,19]. However, the lability and the lack of directionality typical for the dative bonds formed by Ln(III) ions render the diastereoselective synthesis of enantiopure, self-assembled polynuclear structures quite challenging. To date, only a few ligands, such as the ones shown in Figure 1, have been reported to successfully control the structure and to induce the chirality of the resulting architectures. With the introduction of two stereogenic centers on ligand E1, the complexation of Eu(III) leads to the formation of a mixture of a helicate and a tetrahedral cage [20]. Similarly, a heptameric, homochiral Eu(III) wheel could be obtained both directly or stepwise, by using the chiral ligand, E2 [21].
Previously, our group succeeded to functionalize the (–)-5,6-pinene bipyridine ligand by a diastereoselective introduction of a carboxylic arm into the pinene moiety. The ligand obtained, (+)-HL0, was successfully used as a chiral building block for the self-assembly of enantiopure, trinuclear helicates with Eu(III) [22,23]. This complex also showed a solvent-dependent trinuclear-tetranuclear interconversion [24]. In the trinuclear compound [Ln3(L0)63-OH)(H2O)3]2+ there are two sets of ligands, each showing a different coordination mode. Three ligands are bidentate and connect two adjacent metal centers by coordinating solely via their carboxylate group. They form 5-membered chelate cycles. The other three are tetradentate and besides the carboxylate they are coordinating as well with their two nitrogens from the bipyridine unit leading to 6-membered chelate cycles (Figure 2a).
To understand the influence of the ligand design on the self-assembly process and the stability of the obtained architectures, we decided to modify the ligand (+)-HL0 by increasing the length of the carboxylic pendant arm. Thus, we synthesized the new ligands (–)-HL1 (one more CH2 unit) and (–)-HL2 (two more CH2 units) (Figure 2b). Here, we report the results of our study about the influence of the pendant arm length on the coordination behavior with various Ln(III) ions.

2. Materials and Methods

2.1. General

Reagent grade chemical, including the anhydrous solvents, were purchased from Sigma Aldrich and Acros Organics and used without further purification. Eu(ClO4)3nH2O and Lu(ClO4)3nH2O were synthesized as previously reported [25]. Caution! Perchlorate salts are potentially explosive. The amount of water in Ln(ClO4)3nH2O (n = 7.6 for Ln = La, n = 6.3 for Ln = Eu, n = 5.6 for Ln = Lu) was determined by Karl–Fischer titrations on a Mettler Toledo C10s coulometric Karl–Fischer titrator. Yields reported are for isolated, spectroscopically pure compounds. When inert conditions were required, the reactions were performed under the Ar atmosphere using glassware dried overnight at 120 °C. Analytical thin layer chromatography (TLC) was performed on SiO2 plates GF254 (0.25-mm layer thickness). Flash chromatography purifications were performed on a CombiFlash EZ Prep from Teledyne®. Optical rotations were measured on an Anton Paar Modular Circular Polarimeter MCP 100. The measurements were carried out in a quartz vessel (l = 100 mm) with the sodium D line of a sodium lamp (589 nm) using spectroscopic-grade solvents. IR spectra were recorded as solids between 4000–400 cm−1 on a Bruker ALPHA FTIR. UV-Vis spectra were recorded at 20 °C on an Evolution 220 spectrometer from Thermo Scientific equipped with a thermostat. Fluorescence spectra were recorded at 20 °C on a Fluoromax 4 spectrometer from HORIBA equipped with a thermostat. NMR spectra were recorded on a Bruker Advance DPX 300 spectrometer using TMS or the residual solvent proton as internal standard. HRMS spectra were recorded on FTMS 4.7T BioAPEX II and Waters SynaptG2-Si. ESI-MS spectra were recorded on a Thermo Trace DSQ II and Voyager GC-MS. Crystallographic data were collected on a STOE IPDS-II diffractometer.

2.2. General Procedure for (–)-PL

Under inert atmosphere, freshly distilled diisopropylamine (DIA) (5.25 mL, 37.1 mmol) was dissolved in 210 mL of anhydrous THF. The solution was cooled to −20 °C. BuLi (38 mmol 17.5 mL, 2.2 M in hexane) was added. The reaction mixture was heated at 0 °C and left stirring for 10 minutes. Afterwards the mixture was cooled to −40 °C and a solution of (–)-5,6-pinene bipyridine (7 g, 28 mmol) in anhydrous THF (105 mL) was added dropwise for 30 minutes. The solution instantly turned dark blue. The reaction mixture was then stirred for 2 h at −40 °C. After this time, ally bromide (1.5 equiv., 5.08 g, 3.6 mL, 42 mmol) (for (–)-PL1) or 4-bromo-1-butene (1.5 equiv., 5.67 g, 4.3 mL, 42 mmol) (for (–)-PL2) was added and the mixture was stirred at −40 °C for 1 h and then allowed to reach RT. Gradually, the mixture turned dark green and afterwards light brown. The mixture was stirred for 16 h. Afterwards, the remaining base was quenched with 2.1 mL H2O and the solvent was removed under rotary evaporation. The compound was purified by column chromatography on silica 60, eluent: hexane/EtOAc = 4/1.
(–)-PL1: Rf = 0.6. Yield: 76%. 1H NMR (300 MHz, CDCl3) δ 8.64 (ddd, J = 4.8, 1.8, 0.9 Hz, 1H, H1), 8.44 (dt, J = 8.0, 1.1 Hz, 1H, H4), 8.08 (dd, J = 7.8, 0.7 Hz, 1H, H7), 7.80 (ddd, J = 8.0, 7.5, 1.8 Hz, 1H, H3), 7.31 (d, J = 7.8 Hz, 1H, H8), 7.29–7.22 (m, 1H, H2), 5.97 (ddt, J = 16.7, 10.2, 6.5 Hz, 1H, H20), 5.12 (ddt, J = 17.1, 2.2, 1.5 Hz, 1H, H21a), 5.01 (ddt, J = 10.2, 2.3, 1.2 Hz, 1H, H21b), 3.09 (dt, J = 10.0, 3.3 Hz, 1H, H11), 2.80 (t, J = 5.7 Hz, 1H, H14), 2.62–2.53 (m, 1H, H13a), 2.52–2.42 (m, 1H, H12), 2.37–2.27 (m, 2H, H12, H19), 1.72–1.53 (m, 2H, H18), 1.44 (s, 3H, H17), 1.34 (d, J = 9.8 Hz, 1H, H13b), 0.65 (s, 3H, H16). 13C NMR (75 MHz, CDCl3) δ 159.60, 156.89, 153.19, 148.97, 142.12, 139.13, 136.77, 133.54, 123.03, 120.83, 117.75, 114.45, 46.92, 43.74, 43.46, 41.20, 32.07, 31.82, 28.54, 26.45, 20.94. HRMS (ESI) calcd. for C20H23N2+ [M + H]+ 291.1861, found 291.1904. [ a ] D 20 (0.250 g/L in CH2Cl2): −115°.
(–)-PL2: Rf = 0.6. Yield: 83%. 1H NMR (300 MHz, CDCl3) δ 8.64 (ddd, J = 4.8, 1.8, 0.9 Hz, 1H, H1), 8.44 (dt, J = 8.0, 1.1 Hz, 1H, H4), 8.08 (dd, J = 7.8, 0.8 Hz, 1H, H7), 7.88–7.70 (m, 1H, H3), 7.31 (d, J = 7.8 Hz, 1H, H8), 7.28–7.22 (m, 1H, H2), 5.97 (ddt, J = 16.8, 10.2, 6.5 Hz, 1H, H20), 5.17–5.08 (m, 1H, H21a), 5.01 (ddt, J = 10.2, 2.3, 1.2 Hz, 1H, H21b), 3.09 (dt, J = 10.0, 3.3 Hz, 1H, H11), 2.80 (t, J = 5.7 Hz, 1H, H14), 2.64–2.52 (m, 1H, H13a), 2.52–2.39 (m, 1H, H12), 2.39–2.26 (m, 3H, H19), 1.69–1.52 (m, 2H, H18), 1.44 (s, 3H, H17), 1.34 (d, J = 9.8 Hz, 1H, H13b), 0.65 (s, 3H, H16). 13C NMR (75 MHz, CDCl3) δ 159.61, 156.89, 153.19, 148.97, 142.12, 139.13, 136.77, 133.54, 123.03, 120.83, 117.75, 114.45, 46.92, 43.74, 43.46, 41.20, 32.07, 31.82, 28.54, 26.45, 20.94. HRMS (ESI) calcd. for C21H25N2+ [M + H]+ 305.20122, found 305.20091. [ a ] D 20 (0.100 g/L in CH2Cl2): −130°.

2.3. General Procedure for (–)-HL

Over (–)-PL (1.72 mmol) a mixture of acetone (17 mL), water (6 mL) and CH3COOH (8 mL) was added. The suspension was cooled to 0 °C and solid KMnO4 (5 equiv., 17.22 mmol, 2.721 g) was added in 5 portions over a period of 4 h at this temperature. The mixture was brought to RT and stirred for one more hour. After this time, a H2O2 35% solution was added dropwise until the complete dissolution of the precipitate. The obtained solution was extracted with CH2Cl2 (3 × 30 mL). The organic phase was washed with brine and dried over anhydrous MgSO4. The solvent was evaporated, and the crude product was purified by column chromatography on silica gel 60, eluent: CH2Cl2/MeOH = 5/0.1.
(–)-HL1: Obtained: 0.46 g light yellow oil. Yield: 43%. 1H NMR (300 MHz, CDCl3) δ 8.69 (ddd, J = 4.8, 1.8, 0.9 Hz, 1H, H1), 8.23–8.14 (m, 1H, H4), 8.06 (d, J = 7.8 Hz, 1H, H7), 7.82 (td, J = 7.8, 1.8 Hz, 1H, H3), 7.42 (dd, J = 7.9, 1.2 Hz, 1H, H8), 7.30 (ddt, J = 6.9, 4.9, 1.0 Hz, 1H, H2), 3.25 (dd, J = 9.0, 6.6 Hz, 1H, H11), 2.93 (ddd, J = 14.6, 10.0, 7.6 Hz, 1H, H19a), 2.84 (t, J = 5.7 Hz, 1H), H14, 2.75–2.66 (m, 1H, H19b), 2.66–2.56 (m, 1H, H13a), 2.25 (td, J = 6.0, 2.5 Hz, 2H, H18a, H12), 2.00 (ddd, J = 16.9, 13.9, 7.0 Hz, 1H, H18b), 1.44 (s, 3H, H17), 1.38 (d, J = 10.0 Hz, 1H, H13b), 0.64 (s, 3H, H16). 13C NMR (75 MHz, CDCl3) δ 176.89, 158.35, 155.60, 152.85, 149.31, 143.03, 137.27, 134.77, 123.50, 120.98, 119.15, 46.96, 45.47, 42.06, 41.71, 33.63, 29.25, 28.92, 26.20, 20.90. HRMS (ESI) calcd. for C19H21N2O2+ [M+H]+ 309.13975, found 309.15939. [ a ] D 20 (0.552 g/L in CH2Cl2): −109°.
(–)-HL2: Yield: 40%. 1H NMR (300 MHz, CDCl3) δ 8.69 (ddd, J = 4.8, 1.8, 0.9 Hz, 1H, H1), 8.23–8.14 (m, 1H, H4), 8.06 (d, J = 7.8 Hz, 1H, H7), 7.82 (td, J = 7.8, 1.8 Hz, 1H, H3), 7.42 (dd, J = 7.9, 1.2 Hz, 1H, H8), 7.30 (ddt, J = 6.9, 4.9, 1.0 Hz, 1H, H2), 3.25 (dd, J = 9.0, 6.6 Hz, 1H, H11), 2.93 (ddd, J = 14.6, 10.0, 7.6 Hz, 1H, H19a), 2.84 (t, J = 5.7 Hz, 1H, H14), 2.75–2.66 (m, 1H, H19b), 2.66–2.56 (m, 1H, H13a), 2.25 (td, J = 6.0, 2.5 Hz, 1H, H18a), 2.00 (ddd, J = 16.9, 13.9, 7.0 Hz, 1H, H18b), 1.44 (s, 3H, H17), 1.38 (d, J = 10.0 Hz, 1H, H13b), 0.64 (s, 3H, H16). 13C NMR (75 MHz, CDCl3) δ 177.01, 158.47, 155.72, 152.97, 149.43, 143.15, 137.40, 134.89, 123.62, 121.11, 119.27, 47.09, 45.60, 42.19, 41.84, 33.77, 29.39, 29.05, 26.34, 21.03. HRMS (ESI) calcd. for C20H23N2O2+ [M + H]+ 323.1760, found 323.1773. [ a ] D 20 (0.552 g/L in CH2Cl2): −140°.

2.4. General Procedure for the Synthesis of [LnL2](ClO4)

A solution of (–)-HL (0.08 mmol, 1 equiv.) in MeOH (0.35 mL) was treated with N,N-diisopropylethylamine (DIPEA, 14 μL, 0.08 mmol, 1 equiv.) and stirred for 5 min at RT. A saturated solution of Ln(ClO4)3nH2O (0.04 mmol, 0.5 equiv.) in MeOH (0.2 mL) was added in one portion, and 2 mL i-Pr2O were added to the reaction mixture and a white precipitate formed instantly. This was filtered, washed with a cold mixture of MeOH/i-Pr2O (1:1), and dried under a high vacuum.
[La{(–)-L1}2](ClO4)
IR (cm−1): 3112w (νO–H), 2930w (νC–H), 1556s (νas COO), 1428s (νs COO), 1057s (νCl–O), 621s (νCl–O).
(+)-ESI-MS: m/z (%) 309 (100) [(–)-HL1+H]+, 639.9 (23) [2(–)-HL1+Na]+, 752.8 (7) [La{(–)-L1}2]+.
HRMS: calcd. for C38H38N4O4La+ [La{(–)-L1}2]+ 753.1745, found 753.1742.
[Eu{(–)-L1}2](ClO4)
IR (cm−1): 3107w (νO–H), 2913w (νC–H), 1556s (νas COO), 1428s (νs COO), 1045s (νCl–O), 621s (νCl–O).
(+)-ESI-MS: m/z (%) 309 (100) [(–)-HL1+H]+, 766.8 (60) [Eu{(–)-L1}2]+.
HRMS: calcd. for C38H38N4O4Eu+ [Eu{(–)-L1}2]+ 765.1841, found 765.1180.
[Lu{(–)-L1}2](ClO4)
IR (cm−1): 3111w (νO–H), 2918w (νC–H), 1557s (νas COO), 1428s (νs COO), 1093s (νCl–O), 621s (νCl–O).
(+)-ESI-MS m/z (%) 309 (100) [(–)-HL1+H]+, 638.8 (45) [2(–)-HL1+Na]+, 788.8 (20) [Lu{(–)-L1}2]+.
HRMS: calcd. for C38H38N4O4Lu+ [Lu{(–)-L1}2]+ 789.2301, found 789.2283.
[La{(–)-L2}2](ClO4)
IR (cm−1): 2930w (νC–H), 1575s (νas COO), 1430s (νs COO), 1090s (νCl–O), 621s (νCl–O).
(+)-ESI-MS: m/z (%): 323.2 (100) [(–)-HL2+H]+, 781.1 (5) [La{(–)-L2}2]+.
HRMS: calcd. for C40H42N4O4La+ [La{(–)-L2}2]+ 781.2270, found 781.2242.
[Eu{(–)-L2}2](ClO4)
IR (cm−1): 3112w (νO–H) 2930w (νC–H), 1555s (νas COO), 1426s (νs COO), 1096s (νCl–O), 621s (νCl–O).
(+)-ESI-MS m/z (%) 323 (100) [(–)-HL2+H]+, 794.9 (70) [Eu{(–)-L2}2]+.
HRMS: calcd. for C40H42N4O4Eu+ [Eu{(–)-L2}2]+ 793.2193, found 793.2206.
[Lu{(–)-L2}2](ClO4)
IR (cm−1): 3112w (νO–H) 2930w (νC–H), 1579s (νas COO), 1429s (νs COO), 1096s (νCl–O), 621s (νCl–O).
(+)-ESI-MS: m/z (%) 323 (10) [(–)-L2+H]+, 816.9 (100) [Lu{(–)-L2}2]+.
HRMS: calcd. for C40H42N4O4Lu+ [Lu{(–)-L2}2]+ 817.2593, found 817.2614.

2.5. UV-Vis Titrations

Stock solutions of ~0.2 mM [(–)-L][(DIPEA)H] were prepared in anhydrous MeCN (100 mL) by dissolution of [(–)-L][(DIPEA)H] (0.02 mmol: 8.68 mg [(–)-L1][(DIPEA)H] and 9.03 mg [(–)-L2][(DIPEA)H]), which were previously prepared by reaction between (–)-HL and DIPEA in DCM. These solutions were afterwards diluted with anhydrous MeCN to reach a final concentration of ~0.2 mM. The exact concentration of [(–)-L][(DIPEA)H] was determined by UV-Vis spectrophotometry. Stock solutions of Ln(ClO4)3nH2O were prepared in anhydrous MeCN by quantitative dissolution of the salts at a concentration of ~10 mM, followed by dilution until a concentration of ~1 mM. The exact concentration of Ln(III) was determined afterwards by ICP-OES, then 2 mL of ~0.2 mM [(–)-L][(DIPEA)H] solutions were titrated with microvolumes of Ln(III) solutions into UV-Vis cells with a path length of 1 cm, at 20 °C. Each titration was carried out in triplicate and the fitting was done on each titration. The results were averaged to give the final association constants.

2.6. 1H NMR Titrations

Stock solutions of [(–)-L][(DIPEA)H] of 0.167 M concentration were prepared by dissolution in MeOD; 1 M solutions of La(ClO4)3∙7.6H2O were prepared by dissolution of the salt in MeOD; 1 mL of [(–)-L][(DIPEA)H] solutions were titrated with microliters of La(III) solutions into an NMR tube.

2.7. X-Ray Structure Determinations

Single crystals of (–)-PL1, (–)-HL2 were selected and mounted on loop with oil on a STOE IPDS II diffractometer. Crystals were kept at 250(2) K during data collection. A single crystal of (–)-HL1 was selected and mounted on loop with oil on a STOE STADIVARI diffractometer. Crystals were kept at 200(2) K during data collection. Using Olex2 [25], structures were solved with the SHELXT [26] structure solution program using intrinsic phasing and refined with the SHELXL [27] refinement package using least squares minimization.
Crystal data for:
(–)-HL1 C19H20N2O2 (M = 308.37 g/mol): orthorhombic, space group P212121 (no. 19), a = 6.50130(10) Å, b = 10.1527(3) Å, c = 24.2770(5) Å, V = 1602.42(6) Å3, Z = 4, T = 200(2) K, μ(Cu Kα) = 0.668 mm−1, Dcalc = 1.278 g/cm3, 16497 reflections measured (14.102° ≤ 2Θ ≤ 137.64°), 2908 unique (Rint = 0.0181, Rsigma = 0.0125) which were used in all calculations. The final R1 was 0.0317 (I > 2σ(I)) and wR2 was 0.0805 (all data).
(–)-HL2: C40H44N4O4 (M = 644.79 g/mol): monoclinic, space group P21 (no. 4), a = 11.3097(8) Å, b = 6.2574(5) Å, c = 24.2538(19) Å, β = 95.828(6)°, V = 1707.6(2) Å3, Z = 2, T = 250(2) K, μ(MoKα) = 0.082 mm−1, Dcalc = 1.254 g/cm3, 5874 reflections measured (3.376° ≤ 2Θ ≤ 50.516°), 5874 unique (Rint = MERG, Rsigma = 0.0900) which were used in all calculations. The final R1 was 0.0896 (I > 2σ(I)) and wR2 was 0.2532 (all data). Twinned data refinement scales 0.732(6):0.268(6).
(–)-PL1: C20H22N2 (M = 290.39 g/mol): orthorhombic, space group P212121 (no. 19), a = 7.2613(5) Å, b = 9.0056(6) Å, c = 25.277(2) Å, V = 1652.9(2) Å3, Z = 4, T = 250(2) K, μ(MoKα) = 0.068 mm−1, Dcalc = 1.167 g/cm3, 21498 reflections measured (3.222° ≤ 2Θ ≤ 50.25°), 2947 unique (Rint = 0.1080, Rsigma = 0.0646) which were used in all calculations. The final R1 was 0.0400 (I > 2σ(I)) and wR2 was 0.0850 (all data).

3. Results and Discussion

3.1. Synthesis

The ligands (–)-HL1 and (–)-HL2 were obtained by introducing a pendant arm end-functionalized with a carboxylic acid moiety to (–)-5,6-pinene bipyridine [28]. The first step involved the diastereoselective deprotonation of the pinene unit by the bulky base LDA at −40 °C, followed by nucleophilic attack on brominated allylic derivatives, leading to the formation of (–)-PL1 and (–)-PL2 in 76%, respectively, 83% yield (Scheme 1). In the second step, these allylic derivatives were oxidized in the presence of KMnO4, leading to the ligands (–)-HL1 and (–)-HL2, in 43%, respectively 40% yields. The formation of (–)-PL1, (–)-PL2, (–)-HL1, and (–)-HL2 was confirmed in solution by 1H NMR, 13C NMR, and HRMS.
The structures of (–)-PL1, (–)-HL1, and (–)-HL2 were also confirmed in the solid state by single crystal X-ray diffraction.
The structure of (–)-PL1 (Figure 3) confirms the diastereoselective attachment of the pendant arm with the C13 atom in the S configuration, in agreement with previously reported similar reactions on the pinene unit [29]. The compound crystalizes in the P212121 (no. 19) chiral space group, with one molecule in the asymmetric unit.
The compound (–)-HL1 (Figure 4, left) crystallizes in the P212121 (no. 19) space group, with one molecule in the asymmetric unit. Intermolecular H-bonds involving the H of the carboxylic group of one molecule and the pyridine nitrogen N1 of the neighbouring one (O2–H2∙∙∙N1) are present. The strength of this interaction, as indicated by the O2-H2-N1 angle (168°) and the H2∙∙∙N1 distance (1.87 Å), seems rather moderate [30]. The two pyridine rings form an angle of 15° between each other. Unlike the structures presented before, the compound (–)-HL2 possesses two molecules in the asymmetric unit. Similarly to (–)-HL1, each molecule is interacting with a neighboring one through moderate-strength H- bonding, as indicated by the H2A∙∙∙N1 and H4A∙∙∙N3 distances (1.90(2), respectively 1.91(2) Å) and the O2-H2A-N1 and O4-H4A-N3 angles (160(2)°) [30]. The angles between the two planes described by the two pyridine rings are 9.3° for the first molecule (depicted) and 10.6° for the second molecule of the asymmetric unit. As in the case of (–)-PL1 and (–)-HL1, the structure of (–)-HL2 does not show any π-π or C−H∙∙∙ π interactions.

3.2. Complexation Studies

(–)-HL1 and (–)-HL2 were employed as ligands for the synthesis of Ln(III) complexes by using the diamagnetic cations La(III) and Lu(III), and the intermediate-sized, but paramagnetic ion, Eu(III).
Previous studies with the ligand (+)-HL0, showed the formation of chiral supramolecular complexes in 1:2 or 1:3 metal-to-ligand ratios (M:L)[31]. Therefore, these two ratios were also used in the present study. The complexation reactions were carried out in the presence of 2 equiv. of a non-coordinating base, the N,N-diisopropylethylamine (DIPEA), by mixing concentrated methanolic solutions of deprotonated ligand with methanolic solutions of Ln(ClO4)3nH2O salts. The La(III), Eu(III), Lu(III) complexes were isolated by precipitation with diisopropyl ether. The complexes were analyzed by ESI-MS spectrometry, IR, and 1H NMR spectroscopy.
The ESI-MS and IR spectra of the complexes formed with 1:2 and 1:3 M:L ratios, respectively, showed no significant difference, suggesting the formation of similar species in these conditions. In the ESI-MS spectra (see Supplementary Materials) recorded in MeOH, only the peak corresponding to [Ln{(–)-L1}2]+ or [Ln{(–)-L2}2]+ was found. No multinuclear species were detected as was the case of the complexes obtained with (+)-HL0 [22,23,32].
Further, the IR spectra of the amorphous powders were almost superimposable, suggesting isostructurality. Moreover, all showed strong bands at 1086 and 621 cm−1, which stemmed from the stretching vibrations of the uncoordinated perchlorate ion [33]. Another specific feature of the spectra was the presence of a band centered at around 3112 cm−1, which was indicative of O–H stretching. In addition, the two strong bands at 1580 and 1430 cm−1 stemmed from the two vibrational modes of the carboxylate moiety. According to the literature data [34,35,36,37], the coordination mode of the carboxylate unit can be determined from the difference Δν = νasνs of the carboxylate ion. In all the complexes isolated with both ligands, the difference Δν was 150 cm−1. This difference suggests a bidentate–cyclic coordination.
The 1H NMR spectra of the compounds [LaL2](ClO4) (L = (–)-L1, (–)-L2) in MeOD, CD2Cl2, and CDCl3 showed broad signals at RT and also at higher (323 K) or lower temperature (233 and 183 K), indicating fluxionality or possible isomerization (Supplementary Materials, Figures S13–S16).
The Eu(III) complexes were also characterized by emission spectroscopy in CH2Cl2 (λex = 270 nm, Figure 5). The typical narrow transition bands characteristic for Eu(III) complexes dominate both spectra and an efficient antenna effect from the ligands is observed. The presence of an antenna effect further suggests an efficient energy transfer from the ligands towards the metal center, thus indicating a coordination from at least one N atom from the bipyridine moieties. Moreover, the fine structure of the emission bands of both complexes can provide further information regarding the Eu(III) environment ion and its symmetry [15]. While the 5D07F1 transition is independent of the ligand field, the 5D07F2 transition is not, and is particularly influenced by the symmetry around the Eu(III) cation. A decreased symmetry around the metal ion center led to the increase of the 5D07F2 band intensity and thus the 7F2/7F1 intensity ratio could be used to evaluate the coordination environment. Indeed, according to the literature [37], a ratio below 1 was observed for a centrosymmetric environment, while for asymmetric systems, the ratio is typically comprised between 8 and 12. In our case, the ratios are 3.0 for [Eu{(–)-L1}2]ClO4 and 4.8 for [Eu{(–)-L2}2]ClO4, indicating in both cases an asymmetric coordination environment around the Eu(III) ion.

3.3. Determination of Association Constants by UV-Vis Titrations

Further, the association constants of the Ln(III) complexes (Ln = La, Eu, Lu) with the ligands (–)-L1 and (–)-L2 were determined by UV-Vis titrations in MeCN. For this, solutions of the deprotonated ligands (~0.02 mM) were titrated with increments of the corresponding Ln(III) solutions in MeCN (~1 mM). Due to the weak coordinating nature of the perchlorate ion, as confirmed by IR, perchlorate lanthanide salts were employed for the titrations (Ln(ClO4)3nH2O). Previous studies [38] have shown that the trans–cis isomerization (Figure 6) of the ligands can be followed through UV-Vis.
When the ligand is in its free, uncoordinated form, the maximum of absorption is situated at 290 nm. The addition of coordinating metal ions induces a bathochromic shift to 315 nm, which is characteristic for the cis conformation of the bipyridine unit. The evolution of the UV-Vis spectra of the two ligands can be observed in Figure 7.
A global analysis of the spectra was further performed using the open-source titration Fitter [39] software. For the model, only mononuclear species were considered, as shown below:Chemistry 04 00002 i001
The association constants obtained after the numerical fit of the curves are presented in Table 1. The values for the association constants obtained for both (–)-L1 and (–)-L2 were comparable along the series, following for each stoichiometry the expected trend (La3+ < Eu3+ < Lu3+) that reflects the continuous contraction of the ionic radius of lanthanides with increasing atomic number, the concomitant increase in charge density, and thus, the stronger electrostatic metal-ligand interaction. In addition, for both ligands and for any of the three cations, there was a slight decrease in binding constants moving from the species [LnL]2+ towards [LnL3] (i.e., K1 > K2 > K3), in agreement with the steady decrease of (i) the electrical positive charge of the metal species reacting with L, (ii) the number free coordination sites around the lanthanide cation, as expressed by the statistical factors [37], and (iii) the possible steric hindrance from the pinene units. The association constants found for ligand (–)-L1 are much stronger compared with those obtained with the ligand (–)-L2. This reflects the major influence of the size of the chelate ring on the stability of the complexes. In the complexes with (+)-L0, the chelate cycles containing the N from the pinene bipyridine moiety and an O from the carboxylic unit have 6 atoms, while for the (–)-L1 and (–)-L2, 7- and 8-membered rings are formed, respectively. Thus, an increase of the length of the pendant arm disfavours complexation. Other tridentate bipyridine ligands containing the carboxylate ion and forming 5-atom chelate rings, such as those reported by Ziessel and coworkers [40], show cumulative association constants of 5.8(1), 10.6(1), and 14.6, respectively, lower than those obtained with (–)-L1, but similar to the ones obtained for (–)-L2. They also observed higher differences in association constants along the lanthanide series, which is not the case with (–)-L1 and (–)-L2. A thorough comparison is however difficult because the solvent used in both studies are different (water vs. acetonitrile). However, a big chelate cycle is not the only culprit for diminishing the association constants. In the case of an octadentate, bis-phosphonated bipyridine ligand, very high association constants of around 19 were determined for the La(III), Eu(III), and Lu(III) mononuclear complexes, with no trend in association constant along the lanthanide series [41]. The ability of this ligand to enclose the lanthanide ion and shield it from the influence of water molecules is probably the source of the high association constants.
The influence of the size of the chelate ring was further evidenced by the concentration profiles of the formed species (Figure 8). They showed striking differences for the two ligands. In the case of (–)-L2, even in the presence of 2.5 equiv. of La(III) ions, free ligands as well as much lower concentrations of [La{(–)-L2}2]+ and [La{(–)-L2}3] species were observed.

3.4. H NMR Titrations

1H NMR titrations were also carried out on the deprotonated ligands (c~150 mM). Due to jellification in MeCN, the solvent employed was MeOD. La(ClO4)3nH2O was used as a model for the entire series, due to its similar behavior compared to the other Ln(III) ions observed in the UV-Vis titrations, as well as its diamagnetism.
Previous studies [38] have shown that that the trans-cis isomerization of the 5,6-pinene bipyridine can be monitored by 1H NMR. Indeed, the addition of 1 equiv. of acid will isomerize the trans-bpy to its cis form, as the proton is shared between the two N atoms. The addition of 2 equiv. of acid will revert it back to the trans form, with both N atoms being protonated. In our case, the aim of these titrations was to identify the isomerization and the coordination modes of the ligands.
As can be seen in Figures S3 and S4, with increasing increments of La(III), both the aliphatic and aromatic peaks were shifted, but to a different extent. As both (–)-L1 and (–)-L2 have a pendant arm which contains the hard base COO, a good binding group for hard metal ions such as Ln(III), the shifts in the aliphatic regions indicate that this moiety is also involved in complexation. The chemical shifts for the aromatic signal a (Figure 6, Figures S3 and S4) were similar in both cases, indicating the same coordination behavior for both ligands. Firstly, the signal was shielded, reaching a maximum upfield shift Δδ at 0.3 equiv. La(III), while after this point the trend changed and the peak underwent a downfield shift, which, according to the previous studies [38], is indicative of the isomerization. As expected, the aromatic protons (b,c) from the pyridine ring annealed to pinene show negligible shifts. Based on these findings and corroborating with the previously obtained UV-Vis data, we proposed the following binding behavior: First, the ligand binds via the COO group, while the bipyridine unit is still in its trans form, inducing an upfield shift in the unbound trans bipyridine. Afterwards, the bipyridine is starting to bind to La(III), isomerizing from trans to cis, as shown by the downfield of the aromatic protons starting from 0.5 equiv. La(III). The complete isomerization trans to cis is observed at a Ln:L ratio of 1:1.

4. Conclusions

The complexation behavior of two novel chiral ligands, (–)-L1 and (–)-L2, with Ln(III) ions was studied by UV-Vis, emission, and NMR spectroscopy in solution. The formation of three species in equilibrium, namely [LnL]2+, [LnL2]+, and [LnL3], was demonstrated. By direct synthesis, species of the type [LnL2](ClO4) were isolated, as shown by ESI-MS spectrometry and confirmed in the solid state by FT-IR. This behavior is completely different from what it was observed with the parent ligand (+)-L0, which forms discrete, trinuclear species in these conditions. The main explanation is given by the dimensions of the chelate rings, which possess decreasing stabilities in the order: 6 atoms (for (+)-L0) > 7 atoms (for (–)-L1) > 8 atoms (for (–)-L2). Indeed, the association constants confirm that increasing the chelate cycle to 7 atoms (L1) or to 8 atoms (L2) leads to less stable complexes. This study clearly demonstrates how relatively small structural modifications of the ligand can fundamentally change the output of the self-assembly processes.

Supplementary Materials

The following are available online at www.mdpi.com/article/10.3390/chemistry4010002/s1. Figure S1: Superposition of IR spectra of complexes with (–)-L1−; Figure S2: Superposition of IR spectra of complexes with (–)-L2; Figure S3: 1H NMR titration of (–)-L1 with La(III) in MeOD; Figure S4: 1H NMR titration of (–)-L2 with La(III) in MeOD, NMR spectra, (+)-ESI-MS spectra, HRMS spectra. Figure S5: 1H-NMR spectrum of (–)-PL1; Figure S6: 13C-NMR spectrum of (–)-PL1; Figure S7: 1H-NMR spectrum of (–)-HL1; Figure S8: 13C-NMR spectrum of (–)-HL1; Figure S9: 1H-NMR spectrum of (–)-PL2; Figure S10: 13C-NMR spectrum of (–)-PL2; Figure S11: 1H-NMR spectrum of (–)-HL2; Figure S12: 13C-NMR spectrum of (–)-HL2; Figure S13: Stacked 1H-NMR spectra of [La{(–)-L2}2]ClO4 in CD2Cl2 at 298 K (top), 233 K (middle) and 183 K (bottom); Figure S14: Stacked 1H-NMR spectra of [La{(–)-L1}2]ClO4 in MeOD; Figure S15: 1H-NMR spectrum of [Eu{(–)-L1}2]ClO4 in CDCl3; Figure S16: 1H-NMR spectrum of [Lu{(–)-L1}2]ClO4 in CDCl3.

Author Contributions

Conceptualization, O.M. and A.B.S.; methodology, A.B.S. and L.Y.; formal analysis, A.B.S.; investigation, A.B.S., A.C. and L.Y.; resources, O.M., C.A. and K.M.F.; writing—original draft preparation, A.B.S. and O.M.; writing—review and editing, A.B.S., O.M. and C.A.; supervision, O.M. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by HES-SO research funds.

Acknowledgments

The authors wish to thank Albert Ruggi and Fredy Nydegger from University of Fribourg for the HRMS spectra, Felix Fehr for recording the VT 1H NMR spectra and Frédéric Dux for assistance with the titrationFitter software.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Varshey, D.B.; Sander, J.R.G.; Friščić, T.; MacGillivray, L.R. Supramolecular Interactions. In Supramolecular Chemistry; Gale, P.A., Steed, J.W., Eds.; Wiley: Hoboken, NJ, USA, 2012. [Google Scholar]
  2. Kilbas, B.; Mirtschin, S.; Riis-Johannessen, T.; Scopelliti, R.; Severin, K. Dicarboxylate-bridged ruthenium complexes as building blocks for molecular nanostructures. Inorg. Chem. 2012, 51, 5795–5804. [Google Scholar] [CrossRef]
  3. Jansze, S.M.; Cecot, G.; Wise, M.D.; Zhurov, K.O.; Ronson, T.K.; Castilla, A.M.; Finelli, A.; Pattison, P.; Solari, E.; Scopelliti, R.; et al. Ligand Aspect Ratio as a Decisive Factor for the Self-Assembly of Coordination Cages. J. Am. Chem. Soc. 2016, 138, 2046–2054. [Google Scholar] [CrossRef] [PubMed]
  4. Zarra, S.; Wood, D.M.; Roberts, D.A.; Nitschke, J.R. Molecular containers in complex chemical systems. Chem. Soc. Rev. 2015, 44, 419–432. [Google Scholar] [CrossRef] [Green Version]
  5. Metherell, A.J.; Ward, M.D. Stepwise synthesis of a Ru4Cd4 coordination cage using inert and labile subcomponents: Introduction of redox activity at specific sites. Chem. Commun. 2014, 50, 6330–6332. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  6. Von Zelewsky, A.; Mamula, O. The bright future of stereoselective synthesis of co-ordination compounds. J. Chem. Soc. Dalton Trans. 2000, 219–231. [Google Scholar] [CrossRef] [Green Version]
  7. Bünzli, J.C.G.; Piguet, C. Taking advantage of luminescent lanthanide ions. Chem. Soc. Rev. 2005, 34, 1048–1077. [Google Scholar] [CrossRef] [PubMed]
  8. Woodruff, D.N.; Winpenny, R.E.P.; Layfield, R.A. Lanthanide Single-Molecule Magnets. Chem. Rev. 2013, 113, 5110–5148. [Google Scholar] [CrossRef] [PubMed]
  9. Bünzli, J.C.G.; Piguet, C. Lanthanide-containing molecular and supramolecular polymetallic functional assemblies. Chem. Rev. 2002, 102, 1897–1928. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  10. Li, X.Z.; Zhou, L.P.; Yan, L.L.; Dong, Y.M.; Bai, Z.L.; Sun, X.Q.; Diwu, J.; Wang, S.; Bünzli, J.C.; Sun, Q.F. A supramolecular lanthanide separation approach based on multivalent cooperative enhancement of metal ion selectivity. Nat. Commun. 2018, 9, 547. [Google Scholar] [CrossRef] [Green Version]
  11. Lewis, F.W.; Harwood, L.M.; Hudson, M.J.; Drew, M.G.B.; Desreux, J.F.; Vidick, G.; Bouslimani, N.; Modolo, G.; Wilden, A.; Sypula, M.; et al. Highly efficient separation of actinides from lanthanides by a phenanthroline-derived bis-triazine ligand. J. Am. Chem. Soc. 2011, 133, 13093–13102. [Google Scholar] [CrossRef] [Green Version]
  12. Montgomery, C.P.; Murray, B.S.; New, E.J.; Pal, R.; Parker, D. Cell-penetrating metal complex optical probes: Targeted and responsive systems based on lanthanide luminescence. Acc. Chem. Res. 2009, 42, 925–937. [Google Scholar] [CrossRef] [PubMed]
  13. Charbonniére, L.J.; Ziessel, R.; Montalti, M.; Prodi, L.; Zaccheroni, N.; Boehme, C.; Wipff, G. Luminescent lanthanide complexes of a bis-bipyridine-phosphine-oxide ligand as tools for anion detection. J. Am. Chem. Soc. 2002, 124, 7779–7788. [Google Scholar] [CrossRef]
  14. Wei, C.; Ma, L.; Wei, H.; Liu, Z.; Bian, Z.; Huang, C. Advances in luminescent lanthanide complexes and applications. Sci. China Technol. Sci. 2018, 61, 1265–1285. [Google Scholar] [CrossRef]
  15. Klink, S.I.; Grave, L.; Reinhoudt, D.N.; Van Veggel, F.C.J.M.; Werts, M.H.V.; Geurts, F.A.J.; Hofstraat, J.W. A systematic study of the photophysical processes in polydentate triphenylene-functionalized Eu3+, Tb3+, Nd3+, Yb3+, and Er3+ complexes. J. Phys. Chem. A 2000, 104, 5457–5468. [Google Scholar] [CrossRef]
  16. Chow, C.Y.; Eliseeva, S.V.; Trivedi, E.R.; Nguyen, T.N.; Kampf, J.W.; Petoud, S.; Pecoraro, V.L. Ga3+/Ln3+ Metallacrowns: A promising family of highly luminescent lanthanide complexes that covers visible and near-infrared domains. J. Am. Chem. Soc. 2016, 138, 5100–5109. [Google Scholar] [CrossRef]
  17. Barry, D.E.; Kitchen, J.A.; Mercs, L.; Peacock, R.D.; Albrecht, M.; Gunnlaugsson, T. Chiral luminescent lanthanide complexes possessing strong (samarium, Sm III) circularly polarised luminescence (CPL), and their self-assembly into Langmuir–Blodgett films. Dalton Trans. 2019, 48, 11317–11325. [Google Scholar] [CrossRef]
  18. Walton, J.W.; Bari, L.D.; Parker, D.; Pescitelli, G.; Puschmann, H.; Yufit, D.S. Structure, resolution and chiroptical analysis of stable lanthanide complexes of a pyridylphenylphosphinate triazacyclononane ligand. Chem. Commun. 2011, 47, 12289–12291. [Google Scholar] [CrossRef]
  19. Rinehart, J.D.; Long, J.R. Exploiting single-ion anisotropy in the design of f-element single-molecule magnets. Chem. Sci. 2011, 2, 2078. [Google Scholar] [CrossRef]
  20. Li, X.Z.; Zhou, L.P.; Yan, L.L.; Yuan, D.Q.; Lin, C.S.; Sun, Q.F. Evolution of luminescent supramolecular lanthanide M2nL3n complexes from helicates and tetrahedra to cubes. J. Am. Chem. Soc. 2017, 139, 8237–8244. [Google Scholar] [CrossRef]
  21. Bozoklu, G.; Gateau, C.; Imbert, D.; Pécaut, J.; Robeyns, K.; Filinchuk, Y.; Memon, F.; Muller, G.; Mazzanti, M. Metal-controlled diastereoselective self-assembly and circularly polarized luminescence of a chiral heptanuclear europium wheel. J. Am. Chem. Soc. 2012, 134, 8372–8375. [Google Scholar] [CrossRef] [Green Version]
  22. Lama, M.; Mamula, O.; Kottas, G.S.; Rizzo, F.; De Cola, L.; Nakamura, A.; Kuroda, R.; Stoeckli-Evans, H. Lanthanide class of a trinuclear enantiopure helical architecture containing chiral ligands: Synthesis, structure, and properties. Chem. Eur. J. 2007, 13, 7358–7373. [Google Scholar] [CrossRef]
  23. Mamula, O.; Lama, M.; Telfer, S.G.; Nakamura, A.; Kuroda, R.; Stoeckli-Evans, H.; Scopelitti, R. A Trinuclear EuIII Array within a diastereoselectively self-assembled helix formed by chiral bipyridine-carboxylate ligands. Angew. Chem. Int. Ed. 2005, 44, 2527–2531. [Google Scholar] [CrossRef]
  24. Lama, M.; Mamula, O.; Kottas, G.S.; De Cola, L.; Stoeckli-Evans, H.; Shova, S. Enantiopure, supramolecular helices containing three-dimensional tetranuclear Lanthanide (III) arrays: Synthesis, structure, properties, and solvent-driven trinuclear/tetranuclear interconversion. Inorg. Chem. 2008, 47, 8000–8015. [Google Scholar] [CrossRef] [PubMed]
  25. Dolomanov, O.V.; Bourhis, L.J.; Gildea, R.J.; Howard, J.A.K.; Puschmann, H. OLEX2: A complete structure solution, refinement and analysis program. J. Appl. Cryst. 2009, 42, 339–341. [Google Scholar] [CrossRef]
  26. Sheldrick, G.M. SHELXT—Integrated space-group and crystal-structure determination. Acta Cryst. A 2015, 71, 3–8. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  27. Sheldrick, G.M. Crystal structure refinement with SHELXL. Acta Cryst. C 2015, 71, 3–8. [Google Scholar] [CrossRef]
  28. Desreux, J.F. Lanthanide Probes in Life, Chemical and Earth Sciences. Theory and Practice; Bünzli, J.-C.G., Choppin, G.R., Eds.; Elsevier Science: Amsterdam, The Netherlands, 1989. [Google Scholar]
  29. Lötscher, D.; Rupprecht, S.; Collomb, P.; Belser, P.; Viebrock, H.; Von Zelewsky, A.; Burger, P. Stereoselective synthesis of chiral pinene [5,6] bipyridine ligands and their coordination chemistry. Inorg. Chem. 2001, 40, 5675–5681. [Google Scholar] [CrossRef] [PubMed]
  30. Steiner, T. The hydrogen bond in the solid state. Angew. Chem. Int. Ed. 2002, 41, 49–76. [Google Scholar] [CrossRef]
  31. Lama, M.A. Enantiopure Helical (Supra) Molecular Arrays Containing Lanthanides: Design, Synthesis and Properties; EPFL: Lausanne, Switzerland, 2006; pp. Nr. 3708. [Google Scholar]
  32. Mamula, O.; Lama, M.; Stoeckli-Evans, H.; Shova, S. Switchable chiral architectures containing PrIII Ions: An example of solvent-induced adaptive behavior. Angew. Chem. Int. Ed. 2006, 45, 4940–4944. [Google Scholar] [CrossRef]
  33. Lewis, D.L.; Estes, E.D.; Hodgson, D.J. The infrared spectra of coordinated perchlorates. J. Cryst. Mol. Struct. 1975, 5, 67–74. [Google Scholar] [CrossRef]
  34. Socrates, G. Infrared and Raman Characteristic Group Frequencies: Tables and Charts, 3rd ed.; Wiley: Chichester, UK, 2001. [Google Scholar]
  35. Deacon, G. Relationships between the carbon-oxygen stretching frequencies of carboxylato complexes and the type of carboxylate coordination. Coord. Chem. Rev. 1980, 33, 227–250. [Google Scholar] [CrossRef]
  36. Nakamoto, K. Frontmatter. In Infrared and Raman Spectra of Inorganic and Coordination Compounds; Wiley Online Books; John Wiley & Sons, Inc.: Hoboken, NJ, USA, 2008; pp. 1–147. [Google Scholar]
  37. Kirby, A.F.; Richardson, F.S. Detailed analysis of the optical absorption and emission spectra of Eu3+ in the trigonal (C3) Eu (DBM)3·H2O system. J. Phys. Chem. 1983, 87, 2544–2556. [Google Scholar] [CrossRef]
  38. Düggeli, M.; Christen, T.; Von Zelewsky, A. Protonation behaviour of chiral tetradentate polypyridines derived from α-pinene. Chem. Eur. J. 2005, 11, 185–194. [Google Scholar] [CrossRef]
  39. Dux, F. titrationFitter 2020. Available online: https://pypi.org/project/titrationFitter/ (accessed on 15 November 2020).
  40. Comby, S.; Imbert, D.; Chauvin, A.S.; Bünzli, J.C.G.; Charbonnière, L.J.; Ziessel, R.F. Influence of anionic functions on the coordination and photophysical properties of lanthanide (III) complexes with tridentate bipyridines. Inorg. Chem. 2004, 43, 7369–7379. [Google Scholar] [CrossRef] [PubMed]
  41. Souri, N.; Tian, P.; Lecointre, A.; Lemaire, Z.; Chafaa, S.; Strub, J.-M.; Cianferani, S.; Elhabiri, M.; Platas-Iglesias, C.; Charbonniere, L.J. Step by step assembly of polynuclear lanthanide complexes with a phosphonated bipyridine ligand. Inorg. Chem. 2016, 55, 12962–12974. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Examples of chiral ligands used for the diastereoselective self-assembly of polynuclear lanthanide complexes.
Figure 1. Examples of chiral ligands used for the diastereoselective self-assembly of polynuclear lanthanide complexes.
Chemistry 04 00002 g001
Figure 2. (a) Structure of the trinuclear complex formed by (+)-HL0 with Eu(III) ions as determined by single-crystal X-ray diffraction [23]. (b) The structure of the three analogue ligands.
Figure 2. (a) Structure of the trinuclear complex formed by (+)-HL0 with Eu(III) ions as determined by single-crystal X-ray diffraction [23]. (b) The structure of the three analogue ligands.
Chemistry 04 00002 g002
Scheme 1. Synthetic procedure for the synthesis of ligands (–)-HL1 and (–)-HL2.
Scheme 1. Synthetic procedure for the synthesis of ligands (–)-HL1 and (–)-HL2.
Chemistry 04 00002 sch001
Figure 3. Molecular structure of (–)-PL1 determined by single-crystal X-ray diffraction with ellipsoids at 50% probability. The H atoms were omitted for clarity.
Figure 3. Molecular structure of (–)-PL1 determined by single-crystal X-ray diffraction with ellipsoids at 50% probability. The H atoms were omitted for clarity.
Chemistry 04 00002 g003
Figure 4. Molecular structure of (–)-HL1 (left) and one of the molecules in the asymmetric unit of (–)-HL2 (right), determined by single-crystal X-ray diffraction with ellipsoids at 50% probability. The H atoms were omitted for clarity.
Figure 4. Molecular structure of (–)-HL1 (left) and one of the molecules in the asymmetric unit of (–)-HL2 (right), determined by single-crystal X-ray diffraction with ellipsoids at 50% probability. The H atoms were omitted for clarity.
Chemistry 04 00002 g004
Figure 5. Normalized emission spectra of [Eu{(–)-L}2]ClO4 in CH2Cl2 (0.02 mM, λex = 270 nm).
Figure 5. Normalized emission spectra of [Eu{(–)-L}2]ClO4 in CH2Cl2 (0.02 mM, λex = 270 nm).
Chemistry 04 00002 g005
Figure 6. Isomerization of the bipyridine unit upon complexation.
Figure 6. Isomerization of the bipyridine unit upon complexation.
Chemistry 04 00002 g006
Figure 7. UV-Vis spectra of (–)-L1 in MeCN upon additions of La(III) (left), Eu(III) (middle), and Lu(III) (right) from 0 to 5 equivalents (top) and UV-Vis spectra of (–)-L2 in MeCN upon additions of La(III) (left), Eu(III) (middle), and Lu(III) (right) from 0 to 7 equivalents (bottom).
Figure 7. UV-Vis spectra of (–)-L1 in MeCN upon additions of La(III) (left), Eu(III) (middle), and Lu(III) (right) from 0 to 5 equivalents (top) and UV-Vis spectra of (–)-L2 in MeCN upon additions of La(III) (left), Eu(III) (middle), and Lu(III) (right) from 0 to 7 equivalents (bottom).
Chemistry 04 00002 g007
Figure 8. Concentration profiles of the species formed during the titrations of (–)-L1 (left) and (–)-L2 (right) with La(ClO4)3 in MeCN ([L]Tot = 0.02 mM).
Figure 8. Concentration profiles of the species formed during the titrations of (–)-L1 (left) and (–)-L2 (right) with La(ClO4)3 in MeCN ([L]Tot = 0.02 mM).
Chemistry 04 00002 g008
Table 1. Association constants in MeCN of the complexes formed with (–)-L1 and (–)-L2.
Table 1. Association constants in MeCN of the complexes formed with (–)-L1 and (–)-L2.
LLnlog K1log K2log K3log β1log β2log β3
(–)-L1La7.96.35.67.9 ± 0.214.2 ± 0.419.8 ± 0.4
Eu8.06.45.18.0 ± 0.214.4 ± 0.119.5 ± 0.2
Lu8.76.45.08.7 ± 0.115.1 ± 0.120.1 ± 0.3
(–)-L2La5.14.65.35.1 ± 0.29.7 ± 0.215.0 ± 0.4
Eu5.14.75.45.1 ± 0.19.8 ± 0.115.2 ± 0.1
Lu4.94.95.74.9 ± 0.29.8 ± 0.315.5 ± 0.1
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Solea, A.B.; Yang, L.; Crochet, A.; Fromm, K.M.; Allemann, C.; Mamula, O. Complexation Behavior of Pinene–Bipyridine Ligands towards Lanthanides: The Influence of the Carboxylic Arm. Chemistry 2022, 4, 18-30. https://doi.org/10.3390/chemistry4010002

AMA Style

Solea AB, Yang L, Crochet A, Fromm KM, Allemann C, Mamula O. Complexation Behavior of Pinene–Bipyridine Ligands towards Lanthanides: The Influence of the Carboxylic Arm. Chemistry. 2022; 4(1):18-30. https://doi.org/10.3390/chemistry4010002

Chicago/Turabian Style

Solea, Atena B., Liangru Yang, Aurelien Crochet, Katharina M. Fromm, Christophe Allemann, and Olimpia Mamula. 2022. "Complexation Behavior of Pinene–Bipyridine Ligands towards Lanthanides: The Influence of the Carboxylic Arm" Chemistry 4, no. 1: 18-30. https://doi.org/10.3390/chemistry4010002

Article Metrics

Back to TopTop