Next Article in Journal
Plant-Derived Pesticides as an Alternative to Pest Management and Sustainable Agricultural Production: Prospects, Applications and Challenges
Previous Article in Journal
Conformational Analysis of N-Alkyl-N-[2-(diphenylphosphoryl)ethyl]amides of Diphenylphosphorylacetic Acid: Dipole Moments, IR Spectroscopy, DFT Study
Previous Article in Special Issue
Flexo-Ionic Effect of Ionic Liquid Crystal Elastomers
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Luminescent Ln-Ionic Liquids beyond Europium

by
Cláudia C. L. Pereira
1,
José M. Carretas
2,
Bernardo Monteiro
3,* and
João P. Leal
2,*
1
LAQV-REQUIMTE, Department of Chemistry, NOVA School of Science and Technology, Universidade Nova de Lisboa, 2829-516 Caparica, Portugal
2
Centro de Química Estrutural (CQE), Departamento de Engenharia e Ciências Nucleares (DECN), Campus Tecnológico e Nuclear, Instituto Superior Técnico, Universidade de Lisboa, Estrada Nacional 10, 2695-066 Bobadela, Portugal
3
Centro de Química Estrutural (CQE), Departamento de Engenharia Química (DEQ), Campus Tecnológico e Nuclear, Instituto Superior Técnico, Universidade de Lisboa, Estrada Nacional 10, 2695-066 Bobadela, Portugal
*
Authors to whom correspondence should be addressed.
Molecules 2021, 26(16), 4834; https://doi.org/10.3390/molecules26164834
Submission received: 8 June 2021 / Revised: 2 August 2021 / Accepted: 5 August 2021 / Published: 10 August 2021
(This article belongs to the Special Issue New Trends in Ionic Liquids Application)

Abstract

:
Searching in the Web of Knowledge for “ionic liquids” AND “luminescence” AND “lanthanide”, around 260 entries can be found, of which a considerable number refer solely or primarily to europium (90%, ~234). Europium has been deemed the best lanthanide for luminescent applications, mainly due to its efficiency in sensitization, longest decay times, and the ability to use its luminescence spectra to probe the coordination geometry around the metal. The remaining lanthanides can also be of crucial importance due to their different colors, sensitivity, and capability as probes. In this manuscript, we intend to shed some light on the existing published work on the remaining lanthanides. In some cases, they appear in papers with europium, but frequently in a subordinate position, and in fewer cases then the main protagonist of the study. All of them will be assessed and presented in a concise manner; they will be divided into two main categories: lanthanide compounds dissolved in ionic liquids, and lanthanide-based ionic liquids. Finally, some analysis of future trends is carried out highlighting some future promising fields, such as ionogels.

1. Introduction

What is called an ionic liquid (IL) has a very broad definition, comprising multiple substances possessing a wide diversity of structures and properties. An IL consists of both organic and inorganic ions, and may contain more than one cation or anion. Normally, a substance is considered to be an IL if completely composed of ions, with a melting point below 100 °C. Within ILs, there are electrostatic and dispersive interactions at different length scales, leading to a highly anisotropic character. The ions have a large structural diversity which varies from inorganic to organic, simple to complex, including fully or partially ionized acid or base, organic polymeric metal ions, or metalated coordination polymers [1], giving a boundless variety of cation/anion combinations, estimated around the order of 1019 [2].
The first IL (ethylammonium nitrate) was reported in 1914 by Paul Walden, who never expected that ILs would become such an important scientific area a century later. The number of papers on ISI Web published about ILs in the last five years is more than 38,000; this is explained by several factors: ILs are environmentally friendly solvents with properties such as extremely low vapor pressure, low combustibility, excellent thermal stability, and a wide temperature range in their liquid state. The low volatility and combustibility of ILs, along with the possibility of building ionic liquids in which physical and chemical properties can be fine-tuned, has been a reason for their enormous use in recent years [3,4,5]. They have been extensively used as low-environmental-impact solvents in catalysis and separation techniques, as lubricants and additives, as auxiliaries in analytical techniques, as thermal fluid, and as electrolytes [6,7,8,9,10,11,12,13,14]. Additionally, a wide variety of ILs are transparent through the visible and NIR (near-infrared) spectral regions—a key property for optical solvents [15]—and today new ILs are designed for specialized applications in optics and soft luminescent materials [16].
Lanthanides (Ln) are a group of 15 chemical elements, with atomic numbers 57 through 71, all of which have one valence electron in the 5d shell. All of these elements present similar chemical properties, with +3 being the most stable oxidation state, but the +2 and +4 oxidation states are also common [17] (Figure 1).
Most lanthanides present luminescence [18], and this property has been abundantly explored by studying several lanthanide compounds in different chemical environments. The incorporation of metals into ILs opens the potential to combine additional chemical and physical properties in the ILs. In this light, Ln cations are very interesting for their potential to endow ILs with additional features, as Ln compounds are well known for their outstanding properties, including optical, magnetic, or catalytic activities. Focusing on the luminescent properties of the Ln3+ cations, they present unique optical features over other luminophores, such as excellent luminescence efficiency, tunable emission (choice of Ln3+), large pseudo-Stokes shifts, long-lived excited states, and unique monochromaticity [19]. As such, it is not surprising that Ln compounds have been developed for light amplification and generation, security tags, drug delivery, bioimaging, luminescent sensors, and solar energy conversion, among many other applications [20]. The luminescence of Ln3+ ions can arise from intraconfigurational 4f–4f transitions, interconfigurational 4fn→4fn5d1 transitions, and charge–transfer transitions (ligand to metal and metal to ligand). The emission of Ln3+ ions derived from f–f transition covers the NIR (Pr3+, Nd3+, Sm3++, Eu3+, Dy3+, Ho3+, Er3+, Tm3+, Yb3+) through visible (Pr3+, Nd3+, Sm3+, Eu3+, Tb3+, Dy3+, Ho3+, Er3+, Tm3+) to UV (Gd3+) regions.
It has long been known that the molar absorptivity of lanthanide ions is very low (ε = 2–12 M−1cm−1), but this problem can be easily addressed using intramolecular energy transfer processes commonly known as the “antenna effect”. The “antenna effect” is a concept introduced by S. I. Weissman in the early 1940s [21], which refers the use of highly absorptive organic ligands coordinated with Ln ions that function as light harvesters. The mechanism consists of intramolecular energy transfer from excited states of the ligands to the Ln ion, which dramatically enhances Ln’s luminescence, even under low excitation energy. A detailed explanation of the sensitization of Ln ions is beyond the scope of this manuscript, but there are many publications addressing this issue that an interested reader can consult [22,23]. Figure 2 presents a schematic representation of the antenna effect, as well as its Jablonski diagram including other competing energy transitions within the phosphor molecule.
A major problem in Ln emission is luminescence quenching by energy loss through non-radiative relaxation pathways, due to energy transfer to high-energy local-mode vibrations such C–H, N–H or O–H stretching modes in coordinating ligands and solvent molecules. This is particularly relevant for NIR luminescent Ln phosphors due to the narrow energy gap between their lowest excited state and highest ground state levels. Hence, the number of vibration quanta required for their non-radiative deactivation is smaller [24]. For instance, for Er3+ compounds, two activated O–H vibrations are enough to enable a radiationless return to the ground state. In the case of Ho3+-ILs, only one example has been reported to date, by Mudring et al. [2]. Thus, it is not surprising that NIR luminescence is more sensitive to quenching by water than the visible luminescence of Eu3+ [25]. For a better visualization, Figure 3 presents an energy diagram of the NIR luminescent states of Nd3+, Er3+, and Yb3+, with varying numbers of C–H, N–H, and O–H overtones.
In 2010, a review by Eliseeva and Bünzli made a point of the utility of lanthanide luminescence for a variety of functional materials, including “soft” luminescent materials such as liquid crystals, ionic liquids, and ionogels [26]. A review of lanthanides and ionic liquids, published in 2010 by Mudring and Tang [27], addressed various topics such as catalysis and luminescence. In a 2013 review, Feng and Zhang approached hybrid compounds by analyzing various properties including the luminescence of compounds containing Ln ions [28]. Additionally, ionic liquids containing lanthanides (but not exclusively) have been studied for their magnetic properties [29]. Recently, Prodius and Mudring conducted a review discussing the structural and coordination chemistry of ILs derived from rare earth metals, and their practical applications [20].
The combination of Ln with ILs has risen as a powerful platform for potential application in several hi-tech processes, screen display technologies, anticounterfeiting technology, heat-storage materials, in situ imaging, optical sensing, energy harvesting, advanced luminescent coatings, etc. [30,31].
Due to its outstanding emissive properties, europium has received special attention in optical studies based on its luminescence [32,33,34,35,36,37]. In this review, the focus will be on the luminescence of ionic liquids/lanthanides other than europium, to which much less attention has been given.
The application of ILs in lanthanide chemistry has raised much interest in recent years. Their initial use was as a simple solvent, but they have been playing an increasing role not only as solvents, but also as reagents, templates, binders, linkers, or property modifiers. The following subsections will address several of these roles in the combination of lanthanides with ionic liquids.

2. Luminescent Ln Dissolved in Ionic Liquids (Ln@ILs)

In 2004, Binnemans et al. presented ILs as promising solvents for NIR-emitting lanthanide complexes, due to their properties as polar non-coordinating solvents, capable of solubilizing a large number of Ln complexes [25]. They dissolved Nd3+-tosylate, -bromide, -triflate and -sulfonylimide complexes in 1-alkyl-3-methylimidazolium-ILs containing the same anion as the Nd complexes. NIR luminescence spectra of these Nd3+-salts were measured by direct excitation of the metal ion. Furthermore, intense NIR luminescence was observed upon ligand excitation of the Nd complexes with 1,10-phenanthroline or β-diketonate ligands.
Another manuscript, from the same research group, presents highly luminescent anionic Sm3+ β-diketonate and dipicolinate complexes when dissolved in the imidazolium ionic liquid [C6mim][Tf2N] [38]. A judicious choice of the counterion of the Sm3+-complex ensured the solubility of the salts in the ionic liquid. Luminescence spectra were recorded for the complexes dissolved in the imidazolium ionic liquid (Figure 4), and compared with the luminescence of the same complexes in acetonitrile or water, showing that [C6mim][Tf2N] is a better spectroscopic solvent to study Sm3+ luminescence. High-luminescence quantum yields were observed for all of the Sm3+-β-diketonate complexes in IL solutions.
In 2005, Mudring et al. [39] reported the emission spectra of PrI3 and Pr(Tf2N)3 in the ionic liquid [bmpyr][Tf2N]. After excitation to the 3P1 level, remarkable luminescence not only from the 1D2 level, but also from the 3P0 and even from the 3P1 levels, was observed. It is especially noteworthy that both the solutions of PrI3 and Pr(Tf2N)3 in [bmpyr][Tf2N] show emissions from the 3PJ level with unusually high intensities, even at room temperature. It is more common that the non-radiative population of the 1D2 level from the excited 3PJ states is preferred at this temperature.
These papers led to a joint publication of two articles in 2005 and 2006; anhydrous NdI3 and ErI3 were dissolved in carefully dried batches of the [C12mim][Tf2N] ionic liquid [40]. Binnemans et al. observed an intense NIR emission for both the Nd3+ and Er3+ ions, provided by the low water content of the ionic liquid. When the content of water in the solution increased, even if only exposed to atmospheric moisture, a rapid decrease in the luminescence intensity was noticed. In another publication, the authors reported the optical properties of Ln3+ iodides (Ln = Nd, Dy, Tb) in the ionic liquid [C12mim][Tf2N] [41]. As expected, the absence of any C–H, N–H, or O–H high-energy oscillators in the immediate neighborhood of the Ln3+ ions contributes to excellent luminescence properties, because non-radiative decay becomes less likely when compared to typical solvents.
In another 2005 study, Bünzli et al. used the IL 1-dodecyl-3-methylimidazole chloride, [C12mim]Cl, and doped it with 1 mol-% of the Ln ternary complexes [Ln(tta)3(phen)] (Ln = Nd, Er, Yb) [42], obtaining luminescence at room temperature. The spatial arrangement around the lanthanide metal was very similar in both the mesomorphic sample and the parent β-diketonate complex to that in the europium sample. Moreover, the mesomorphic samples containing Nd, Er, and Yb showed relatively intense NIR luminescence.
Ionic liquids were used by Binnemans et al. [43] as solvents for dispersing luminescent lanthanide-doped LaF3:Nd3+ nanocrystals with a stabilizing ligand (betaine = N,N,N-trimethylglycine). LaF3:Nd3+:betaine could successfully be dispersed in [C4mpyr][Tf2N], [C4mpyr][TfO], and [C4mim][Tf2N]. NIR luminescence was observed for the Nd3+-based systems.
Highly luminescent anionic Sm3+–diketonate and –dipicolinate complexes were dissolved in the imidazolium ionic liquid [C6mim][Tf2N]. The Sm3+ complexes considered by Binnemans et al. [38] were [C6mim][Sm(tta)4], [C6mim][Sm(nta)4], [C6mim][Sm(hfa)4], and [choline]3[Sm(dpa)3]. Luminescence spectra were recorded for the Sm3+ complexes dissolved in the imidazolium ionic liquid, as well as in a conventional solvent. These experiments demonstrate that [C6mim][Tf2N] is a suitable spectroscopic solvent for studying Sm3+ luminescence. High-luminescence quantum yields were observed for the Sm3+–diketonate complexes in solution.
These years provide a series of studies by several research groups. Gatsis and Mudring [44] reported on C12mimBr, a well-known ionic liquid crystal, doped with SmBr3, TbBr3, and DyBr3, which allowed them to obtain ionic liquid crystal materials that show luminescence in the three basic colors (red, green, and orange). Most interestingly, the emission color for the TbBr3- and DyBr3-containing materials can be tuned from bluish white (mainly [C12mim][Br] emission) to green (for TbBr3) or orange-yellow (for DyBr3), depending on the wavelength of the excitation light used.
The luminescence properties of TbCl3(phen)2(H2O)3 in the solid state and in solutions of the [C12mim][Cl] ionic liquid were investigated by Puntus et al. [45]. Luminescence data contributed to elucidating the structural peculiarities of lanthanide chlorides, and revealed a highly efficient luminescence for the Tb3+ complex.
Luminescent soft materials were obtained by Huarong et al. [46] by dissolution of lanthanide (Nd, Er) oxides and organic ligands (tta, phen) into carboxyl-functionalized ionic liquids (IL1 = 3-(5-carboxypropyl)-1-methylimidazolium bromide; IL2 = 3-(5-carboxy propyl)-1-butylimidazolium bromide). Optical properties of the soft materials, such as color and luminescence, can be adapted by simply changing the type of lanthanide ions and/or addition of an organic ligand. The obtained ionic liquids (Nd-tta/IL2, Er-tta/IL-2) present luminescence in the NIR region, and their excitation and emission spectra are very similar to those of Nd-tta/IL-1 and Er-tta/IL-1.
The luminescent properties of Tb3+ dissolved in ionic liquids were studied by Hopkins and Goldey [47]; this study included a simple lanthanide compound (TbCl3) dissolved in a [C4mim][Br]/water mixture, and showed that the [C4mim][Br] ionic liquid does sensitize Tb3+ luminescence.
The emission properties, including luminescence lifetimes, of the lanthanide complexes Ln(Tf2N)3 (Ln3+ = Pr, Nd, Sm, Dy, Er, Tm) in the ionic liquid [bmpyr][Tf2N] were presented by Brandner et al. [48]. The luminescence lifetimes in these systems are remarkably long compared to values typically reported for Ln3+ complexes in solution, reflecting weak vibrational quenching.
1-Butyl-3-methylimidazolium benzoate-IL, [C4mim][BA] was found by Viswanathan et al. [49] to enhance the fluorescence of Tb3+; this enhancement resulted from a sensitization of the lanthanide fluorescence by the benzoate anion of the IL, and a reduction in the non-radiative channels provided by the non-aqueous environment caused by the NIR. The authors also found that the fluorescence enhancement of the lanthanides in the IL was limited due to an inner filter effect, which resulted from strong benzoate absorption. For Tb3+, the strong emission of the ionic liquid in the region 450–580 nm masked the lanthanide emission. To observe the long-lived Tb3+ emission and distinguish it from the short-lived emission from the IL, an appropriate delay was used in the detection.
Mudring et al. [50] report the synthesis of interesting new materials such [C12mim][Br] and [C12mpyr][Br], both doped with TbBr3, as these are able to form mesophases over a wide temperature range. All materials show strong green luminescence from the 5D4 level of Tb3+ after excitation into the 4f8 → 4f75d1 transition. In the case of the imidazolium compound, the color of this emission can be switched between green and blue-white depending on the excitation energy. After excitation with λex = 254 nm, strong green emission is observed—mainly from the 5D4-level of Tb3+—while with λex = 366 nm, only the blue-white luminescence from the imidazolium cation itself is detected.
Li et al. [51] reported the synthesis of carboxylic-acid-functionalized ILs with linear alkyl chains of various lengths on the cations ([Carb-Cnmim]Br, n = 8, 12, 16); they also tested the solubilizing capability of the synthesized ILs relative to Tb4O7, which led to soft luminescent materials combining Ln3+ ions and ILs. The luminescent properties of the obtained materials were investigated.
In 2014, Bortoluzzi et al. [52] prepared the ionic liquid [P8,8,8,1][BrMA] via the addition of HBrMA to [P8,8,8,1][CH3OCO2]. The doped ionic liquid Tb@[P8,8,8,1][BrMA] was obtained by adding anhydrous TbCl3 to [P8,8,8,1][BrMA]. In addition to the measurements performed on the pure complex, the photoluminescence of Tb@[P8,8,8,1][BrMA] was investigated, but the emissions from the metal ions were almost completely masked by an intense and broad band covering the range 400–750 nm, which can be mainly attributed to the fluorescence of the ionic liquid.
Highly luminescent tetrakis Sm3+ complexes with the dbm ligand and the phosphonium [P8,8,8,1]+ counterion were synthesized by Malba et al. [53]. Crystal data from [Sm(dbm)4][P8,8,8,1] show that the Sm3+ ions are surrounded by four dbm ligands coordinating, as expected, in a bidentate fashion, efficiently shielding the metal center from solvent molecules. The photoluminescence of both complexes was studied in the solid state, [P8,8,8,1][Tf2N]-IL, and acetonitrile; as expected, they presented broadening of the photoluminescence emission peaks when passing from the solid state to the complex dissolved in either the IL or a molecular solvent, due to collisions and electrostatic interactions with solvent molecules. The emission level was 4G5/2 for all Sm3+ emissions. The polarizability of the complex determined by the ratio between the integrated areas of the 4G5/26H9/2 and 4G5/26H5/2 transitions, and of the emission spectrum in solid state, was 13.3, which is quite high for a Sm3+ complex, and was similar to that of Eu3+–β-diketonate complexes. Time-resolved analysis of [Sm(dbm)4][P8,8,8,1] presented lifetime values of 63.5 μs, 19.1 μs, and 3.1 μs in the solid state, [P8,8,8,1][ Tf2N]-IL, and acetonitrile, respectively. Internal quantum efficiencies of ~2% in solid state, 1% in IL, and 0.1% in acetonitrile indicate an important contribution of non-radiative recombination pathways in solution.
Ln3+-doped ionic liquids were prepared by dissolving the complex Tb(pybox)3 into a bidentate organophosphine-functionalized ionic liquid (1,3-bis-[3-(diphenylphosphinyl)propyl]imidazole bis(trifluoromethylsulfonyl)imide) by Li et al. [54]. These materials show improved luminescence efficiency, attributed to the coordination of ionic liquids with Ln3+ ions, and can be beneficial for enhancing the photovoltaic energy conversion efficiency of silicon-based solar cells. The authors prepared large-area (17 × 17 cm2) flexible, transparent, luminescent poly(methyl methacrylate) thin films (Figure 5), and applied them as luminescent coatings to the silicon-based heterojunction solar cells, obtaining—in the best case, using the Tb3+-containing film—an increase in performance of ~16%.
Recently, Tang et al. synthesized complexes of La, Nd, Eu, Tb, Dy, and Yb with dicyanamide (DCA) ions—[C2mim][Ln(DCA)4(H2O)4]—using a DCA-based ionic liquid [55]. Luminescence studies were conducted with Eu, Tb, and Dy compounds at room (RT) and liquid nitrogen (LT) temperatures, with the spectra presenting the corresponding characteristic f−f transitions. The emission spectra of the Tb compound were similar to the analogous spectra of the Tb[N(CN)2]3 and Tb[N(CN)2]3·2H2O, although with different relative intensities, which is an indication of different local symmetries around the Tb3+ center. The emission lifetimes (5D4) increased from 0.60 ms at RT to 0.71 ms at LT. In the case of the Dy compound, hypersensitive transition (ΔL = 2, ΔJ = 2) 4F9/26H13/2 was the most intense and, thus, responsible for the yellowish luminescence. The Dy3+ (4F9/2) decay time of 11.9 μs is comparable to that of the series [C6mim]5−x[Dy(SCN)8−x(H2O)x] (x = 0−2) [56], which will be discussed in the Ln-ILs section. The CIE coordinates (chromaticity coordinates—CIE color space is a quantitative link between distributions of wavelengths in the electromagnetic visible spectrum and physiologically perceived colors in human color vision) of the Ln compounds were determined from the respective RT and LT emission spectra, showing a great variety of colors from the red (Eu) to the green-yellow (Tb) and blue-yellow (Dy) regions (Figure 6).
Table 1 presents a list of the Ln@ILs discussed in this review, ordered by Ln3+ center along with their lifetimes and associated excited level and, when available, absolute quantum yield.

3. Luminescent Ln-based Ionic Liquids (Ln-ILs)

The dissolution of Ln salts within ILs is a neat and simple method, but when the anionic moieties of the ILs have no coordinating capability, this method only enables low concentrations of Ln ions due to the low solubility of the salts. However, when an IL with an anionic part has some coordinating ability, higher concentrations can be achieved due the formation of Ln-based ionic liquids (Ln-ILs), in which the Ln ions become part of the ILs in the form of an anionic complex.
In 2003, Jensen et al. studied the stoichiometry of Eu and Nd complexes with the Htta ligand in a biphasic aqueous/[C4mim][Tf2N] system [57]. All of the characterization techniques supported the formation of anionic [Ln(tta)4]- species with no water coordinated to the metal center in the [C4mim][Tf2N]-RTIL phase, instead of the hydrated neutral complexes—Ln(tta)3(H2O)n—that form in the nonpolar molecular solvents xylene or chloroform. The presence of anionic Ln complexes in the IL is made possible by the exchange of [Tf2N]- anions into the aqueous phase in exchange for the [Ln(tta)4]- complex. Additionally, it was shown that the resulting [C4mim][Ln(tta)4] ion pairs exert little influence on the structure of the ionic liquid phase.
Although Ln-ILs can be formed by the dissolution of Ln salts in coordinating ILs, another strategy is to prepare them from the start.
The first Ln-IL series was published in 2006 by Nockemann et al. [58], based on the [C2mim]+ cation and lanthanide thiocyanate anions. The Ln-ionic liquids, prepared via a metathesis procedure, presented the general formula [C2mim]x-3[Ln(NCS)x(H2O)y] (x = 6, Y = 2 (Y); x = 7, Y = 1 (La, Pr, Nd, Sm, Gd, Tb, Ho, Er, and Yb; x = 8, y = 0 (La)), and the luminescence properties of the Sm-ILs were studied a few years later by Ohaion et al. [59]. In this study, the plot of the luminescence intensities of Sm3+ solutions vs. NCS/Ln ratios showed that increasing the NSC/Ln ratio leads to detaching of water molecules from the Sm3+ center, with their replacement by thiocyanate ligands within the coordination sphere. This result supported the composition proposed by the authors—[C2mim]x-3[Ln(NCS)x(H2O)8-x]—which assumed an absence of water for NSC/Ln ratios above 8.
The first studies concerning the emission of Ln-ILs were reported by Mudring et al. in two consecutive manuscripts in 2008. The first publication concerned the low-melting-point Eu-ILs [R]x[Eu(Tf2N)3+x] (x = 1 for R = C3mim and C4mim; x = 2 for C4mpyr) [34], while the second concerned the first examples of room temperature ILs (RTILs) that combine magnetic and luminescent properties, by describing the magneto-optical properties of the [C6mim]5-x[Dy(SCN)8-x(H2O)x](x = 0–2) compounds [56]. All three orange-colored Dy-ILs presented a strong response to a commercial neodymium magnet, and an intense yellow emission characteristic of the Dy3+ center. The most intense transition, between the 4F9/2 and 6H13/2 levels, presented an extremely sharp line shape, indicating high color purity. For the three Dy-ILs, only one Dy3+ is present in the anhydrous [C6mim]5[Dy(SCN)8] analog presenting the highest decay time, explained by the fact that the thiocyanate ligands, unlike water molecules, are not prone to take up the energy of the excited state, thus providing a fairly rigid ligand environment.
In the following year, Getsis et al. reported Dy-based ionic liquid crystals based on the [C12mim]3[DyBr6] compound [60]. This Dy-IL showed interesting luminescent properties together with mesomorphic and superparamagnetic behavior. The [C12mim]3[DyBr6] presented either a bright white or orange yellow emission, depending on the chosen wavelength of excitation. Irradiation with a wavelength of λex = 366 nm leads to a bluish-white luminescence characteristic of the imidazolium moiety, while upon irradiation with a wavelength of λex = 254 nm, the compound turns orange due to the Dy3+ ion emission. By comparing the luminescence spectra of the [C12mim]3[DyBr6] and pure [C12mim]Br, it becomes evident that the bluish-white appearance of the Dy-IL comes from a combination of the luminescence of the [C12mim]+ cations and a small contribution from the Dy3+ transition around 480 nm. It is also interesting that the lifetime of the emission of the most intense transition (4F9/26H13/2) was unaffected by temperature or by the physical state of the compound.
The same group extended these studies to the Tb3+ analog and the [C12mpyr]- anion [51]. To do so, they prepared two new Tb-ILs—[C12mim]3[TbBr6] and [C12mpyr]3[TbBr6]—as well as two samples consisting of [C12mim]Br and [C12mpyr]Br ILs, both doped with TbBr3. All four samples present mesomorphic behavior, and are capable of forming smectic liquid crystalline phases. The Tb-doped ILs crystallize around room temperature, while the neat Tb-ILs solidify as liquid crystal glasses around –5 °C. All compounds present strong green luminescence with the typical Tb3+ emission bands, with long lifetimes of the excited 5D4 level. The pyrrolidinium compounds had somewhat higher lifetimes in comparison with the imidazolium analogs, with the latter presenting a more intense luminescence emission due to an energy transfer from the imidazolium cation to the Tb3+ ion, as was reported in the case of [C12mim]3[DyBr6], as described above [60]. Additionally, just like in the case of the Dy-IL with the [C12mim]- anion, it is possible to tune the color emission of the Tb samples with this imidazolium between green and white by using a UV light excitation of 254 nm or 366 nm, respectively.
Li et al. reported a series of four multifunctional [Dy(SCN)8]-ILs using different phosphonium cations with luminescence, paramagnetism, and tumor mitochondrial targeting properties: [Ph4P]5[Dy(SCN)8], [Ph3PBnOEt]5[Dy(SCN)8], [Ph3PBnNO2]5[Dy(SCN)8], and [Ph3PBn]5[Dy(SCN)8] (Bn = benzyl group) [61]. These Dy-ILs were studied as fluorescence imaging markers in vital cell cultures via confocal laser microscopy (Figure 7). It was found that the uptake of these lipophilic Dy-ILs occurred in the cell membrane, and with selective inhibition of the growth of tumor cells.
In 2015, Tang et al. reported five hexanitratosamarate(III) salts, of which four were Sm-ILs—[Cnmim]3[Sm(NO3)6] (n = 2, 4, 6, 8) [62]. The fifth compound—[MC1mim]3[Sm(NO3)6] (MC1mim = 1,2,3-trimethylimidazolium)—although not an IL, was useful for the structural elucidation of the hexanitratosamarate(III) anion. The three Sm3+ excitation states 4G7/2, 4F3/2, and 4G5/2 could be excited efficiently, but only the emissions from the 4G5/2 level were observed intensively. A possible explanation is that the two higher excitation states 4G7/2 and 4F3/2 are close in energy to the 4G5/2 level, allowing electrons to relax to the lower level quickly through non-radiative transition processes. The lifetime luminescence of all compounds falls in the 114.4–130.3-μs range in acetonitrile, which is longer than that of most Sm3+ compounds. The reason for these high values is the absence of C–H, O–H, or N–H bonds in the [Sm(NO3)6]3−, which would increase non-radiative transitions. In fact, for the [C6mim]3[Sm(NO3)6]-IL, the addition of only 50 μL of water to 5 mL of the Sm-IL led to a sharp decrease in the lifetime from 114.8 μs to 3.75 μs.
Han et al. prepared a series of [C4mim]3[LnCl6] (Ln = Sm, Dy, Er, Yb) crystals from solutions of LnCl3 dissolved in [C4mim][Cl] [63]. Additionally, to study the importance of cross-relaxation within the Sm3+ and Dy3+ samples, they also prepared two samples of these ions diluted in Gd3+ (5% Sm and 5% Dy). From the crystal data, the authors concluded that the first coordination sphere of the Ln3+ ions is composed of six Cl anions, in a slightly distorted octahedral LnCl63– fashion, while the second coordination sphere consists of nine [C4mim]+ cations.
The emission spectra and luminescence lifetimes of both [C4mim]3[LnCl6] crystals and LnCl3 in [C4mim][Cl] solutions were determined in order to study the surroundings of the metals in solution (Figure 8). The spectroscopic similarity found between both spectra suggest that crystalline [C4mim]3[LnCl6] is a good model of the Ln3+ coordination environment in [C4mim][Cl] solutions.
In this study, it deserves to be highlighted that, for this system, the second-coordination-sphere quenching is relatively efficient. For example, for the small-energy-gap Ln3+, the multiphonon relaxation of Ln(Tf2N)3 complexes in [bmpyr][Tf2N] is much less effective [48], notwithstanding the fact that the maximum vibrational energy within the [Ln(Tf2N)x]3−x is ∼1340 cm−1 whilst the phonon cutoff for the LnCl63- is ∼260 cm−1. The explanation for this is that the small radius of the first coordination sphere of the [LnCl6]3− anions does not provide adequate protection from the high-energy C−H oscillators from the [C4mim]+ counterions of the second coordination sphere.
Two new lanthanide-based RTILs—[C4mim][Ln(NO3)4] (Ln = Dy, Sm)—were synthesized and characterized by Fan et al. [64]. The photoluminescence properties of these hydrostable and ecofriendly Ln-ILs were studied at room temperature in deionized water, and their strong fluorescence indicates that these ILs could be used as good luminescent materials. As such, the authors studied their application as fluorescent sensors for Fe(III) (Figure 9).
Both Ln-ILs presented high specific recognition for aqueous Fe3+ ions, even discernible with the naked eye, with no interference by many other common metal ions.
In 2016, Pohako-Esko et al. prepared a series of hexahalocerate(III) salts with the general formula [cation]3[CeHal6] by dissolving anhydrous cerium trihalides in imidazolium halide ionic liquids [65]. Complexes with different halides—[CeCl6]3− and [CeBr6]3—were combined with [C4mim]+, [C6mim]+, [C8mim]+, and [C10mim]+ counterions along with the mixed-halide compound [C4mim]3[CeBr3Cl3]. The melting points of the synthesized salts varied between 85 °C for [C10mim][CeBr6] and 155 °C for [C6mim][CeCl6]. The melting points increase slightly from carbon number n = 4 in the alkyl chain to n = 6, and then continuously decrease with increasing alkyl chain lengths. However, only two of the salts can be considered ILs—[C10mim][CeBr6] and [C4mim]3[CeBr3Cl3]. [C4mim][CeHal6] salts presented intense photoluminescence ascribed to Ce3+-based 5d–4f-centered emission ranging from soft UV to the border of visible emission. The emission was made possible either by direct excitation of the Ce3+ center or by the sensitizing effect of the [C4mim]+ counterion. These results show that emissive Ln-ILs can be designed with parity-enabled Ce3+-based luminescence.
Alvarez-Vicente et al. prepared and characterized the [P66614][LnCl6] and [P4444][LnCl6] of the entire Ln3+ series plus Y3+ and Sc3+, as well as the [P4448][LnCl6] analogs for Ln = Ce, Nd, Sm, Tb, Dy, and Er [66]. The entire [P66614][LnCl6] series, with the longer alkyl chains in the phosphonium, is composed exclusively of room temperature ionic liquids, as commonly found for this cation [67,68,69,70], with melting points (m.p.) between −58 and −40 °C, with the exception of the La3+-IL at –1.6 °C. In the case of the [P4444][LnCl6] series, all compounds are ILs with an m.p. in the range 43–96 °C—again, with the exception of the La3+-IL, which has an m.p. of 103 °C. The [P4448][LnCl6] series showed more irregular behavior and trends than the other series. The lighter lanthanides had an m.p. only slightly lower than the [P4444]+-analogs; however, upon cooling, these compounds behave like supercooled liquids with rather low crystallization temperatures (between 12 and 18 °C). In the case of the heavier lanthanides, the compounds have relatively low melting points, ranging from −6 to −48 °C. EXAFS measurements with selected samples confirmed the LnCl6 coordination in the liquid state, with the Ln···Cl distance decreasing with decreasing Ln ionic radius. The authors performed magnetic, electrochemical, and luminescence studies on selected samples. Concerning the photoluminescence studies, detailed studies were conducted concerning the excitation, emission, and decay times of the visible-light-emitting Tb-, Dy-, and Sm-ILs, as well as the visible-light- and NIR-emitting Nd- and Er-ILs. All transitions in the excitation and emission spectra were assigned, and it is worth highlighting the relatively long luminescence lifetimes of 2.470 μs and 2.725 μs for the [P66614][NdCl6] and [P66614][ErCl6], respectively.
Zheng et al. synthesized novel fluorescent RTILs based on Dy3+ ([MOEmim][Dy(NO3)4]) [71]; they exhibited good fluorescence properties with light blue (Dy3+) luminescence. This is the first time that the application of Ln-RTILs as fluorescence sensors for aromatic compounds has been studied. The sensor worked via the fluorescence-quenching of the phosphor when in the presence of trace amounts of o-(m-, p-)nitrotoluene. Among the three isomeric nitrotoluenes, p-nitrotoluene showed the most significant fluorescence-quenching effect. Furthermore, the two fluorescent ionic liquids demonstrated high selectivity toward nitrotoluene even in the presence of methylbenzene, phenol, chlorobenzene, and aminobenzene. Hence, the selective recognition of nitrotoluene from other aromatic compounds may be used for the analytical detection of explosives.
When designing ILs, highly charged ions are, by default, ruled out in order to avoid high Coulombic attraction that could easily lead to the compound being solid at room temperature (or below 100 °C). However, in 2017, Prodius et al. published the series [Ln5(C2H5-C3H3N2-CH2COO)16(H2O)8](Tf2N)15 (Ln = Er, Ho, Tm; C2H5-C3H3N2-CH2COO = 1-carboxymethyl-3-ethylimidazolium), featuring the pentanuclear Ln-containing +15 cation [Ln5(C2H5-C3H3N2-CH2COO)16(H2O)8]15+ (Figure 10) [2].
These ILs were prepared from the reaction of the respective Ln oxide with 1-carboxymethyl-3-ethylimidazolium chloride and LiTf2N in water, with the ILs forming a separate phase. These ionic liquids show a low tendency to crystallize, taking a few months for the Er-ILs and Ho-IL to form crystals under ambient conditions. In the case of the Tm-IL, all crystallization attempts failed, and only glass transitions were observed. NIR luminescence studies performed at room temperature showed a broad band at 1540 nm for the Er-IL, with four shoulders assigned to the 4I13/24I15/2 transition of the Er3+ ion with an emission lifetime of 0.6 μs. In the case of the Ho-IL, the emission spectra presented a set of bands assign to the 5F55I7, 5I65I8 and 5F55I6 transitions of the Ho3+ ion with an emission lifetime of 0.8 μs. The values of the lifetimes are appreciably high for compounds in the liquid state. Usually, Ln3+ ions surrounded by water molecules and organic ligands present lifetimes in the ns range. These lifetimes are similar to those found for glasses and complexes surrounded by rigid ligands. Additionally, these Ln-ILs presented the highest ever reported values of the effective moments for magnetic ionic liquids, and extraordinary catalytic activity in the three-component synthesis of ethyl 2-methyl-4-(2-oxo-2,3-dihydro-1H-3-indolyl)-5-phenyl-1H-3-pyrrolecarboxylate.
One year later, Cheng et al. reported a series of anhydrous fluorescigenic magnetofluids based on Ln3+ ions [RCnmim]2[Ln(NO3)5] (Ln = Gd, Tb, Dy; R = H or methyl and n = 2, 4, 6, 8) (Figure 11); all of these compounds are RTILs except for R = methyl with n = 1 [30]. The coordination environment composed of nitrate ligands weakened the Ln3+−Ln3+ energy transition, thus avoiding concentration-quenching effects. Additionally, there was no relevant energy exchange between the imidazolium counterions and the Ln3+ centers. The lifetime value of the Tb-IL series varied between 1.262 and 1.294 ms, while the lifetimes of the Dy-ILs varied from 54.75 to 59.27 μs. The fluctuation of the lifetime values within the different series—i.e., with different imidazolium cations—was explained by the interplay of the influences of the different spacing of the anions. The long luminescence lifetimes of the Ln-ILs confirm that these possess a favorable environment that can protect the Ln3+ phosphors from any interference, provided that the structures are integrated.
In 2019, Ramos et al. reported the synthesis of the new organic ionic liquid 1,4-methylimidazolylbutane sulfonate bromide (IL-1), and used it to prepare Ln-ILs Ln(IL-1)3(H2O)3 (Ln = Gd, Eu, Tb) (Figure 12) [31]. The triplet level of the organic IL is ~25,000 cm−1, and the singlet is ~30,000 cm−1. The first excitation state of the Gd3+ ion (6P7/2) is ~32,000 cm−1 so, as expected, the Gd-IL only showed ligand photoluminescence. As for the Tb3+- and Eu3+-ILs, the ligand proved to be more efficient in the sensitization of the Tb3+. Furthermore, the absence of a broad phosphorescence emission band in the range 400–600 nm arising from the ligand indicates that the intramolecular energy transfer from the ligand to the Tb3+ center was very efficient. The lifetimes determined for the Tb-IL were 0.7712 and 0.6988 ms, for direct Tb3+ excitation (370 nm) and ligand excitation (350 nm), respectively. These long lifetime values were the highest reported amongst all Tb soft materials reported.
Wu and Shen, in 2019, reported a new type of magnetic ionic liquid (MIL), incorporating the same Ln ion in both the cationic and anionic parts of the Ln-IL, with the general formula [Ln(TODGA)3][Ln(hfa)4]3 (Ln = Tb, Dy, Ho, Er, Tm, Yb) [72]. It is worth noting that these are the only examples of Ln-ILs with Ln-based cations. Although this was a report focusing on the magnetic properties of the Ln-ILs, the authors described the Tb-IL’s luminescence properties. Upon excitation at 352-nm UV light, the [Tb(TODGA)3][Tb(hfa)4]3 ionic liquid presents the common green luminescence of the Tb3+ ions. By comparison with the luminescence spectra of the isolated cationic and anionic components—Tb(TODGA)3Cl3 and NH4[Tb(hfa)4] (Figure 13)—where the Tb(TODGA)3Cl3 presented very low emission intensity, the Tb-IL presented the highest luminescence emission. As such, the authors concluded that the luminescence emission of the Tb-IL is mainly derived from the anionic part, and that the hfa ligand is more effective than TODGA in the sensitization of Tb3+ ions.
Table 2 presents a list of the Ln-ILs discussed in this review, ordered by Ln3+ center along with their lifetimes, associated excited levels, and physical states/transitions.

4. Overlook and Future Perspectives

To obtain highly efficient Ln molecular light-conversion devices, it is necessary to optimize several parameters: avoid self-quenching channels, use chromophores with high molar absorbance and ideal energy positions of singlet and triplet states (for an efficient energy transfer to Ln3+ ions), while avoiding competitive non-radiative pathways such as multiphonon relaxation to high-energy vibrations (e.g., O–H, C–H, and N–H stretching modes). With this context in mind, the combination of lanthanides and ionic liquids began by using ILs as matrices to protect Ln3+ ions from vibration-induced deactivation processes—mainly from the ever-present water adsorbed in organic solvents. Although ILs proved to provide good protection against the presence of water within the first and/second coordination spheres of the Ln centers, due to the low solubility of the Ln salts, this method usually enabled low concentrations of Ln ions, although higher concentrations could be achieved by the use of the same anionic moieties as both the ligand (of the Ln complex) and the anion (of the ILs). This shortcoming was circumvented by preparing Ln-based ionic liquids, either via direct preparation by metathesis, or by dissolving Ln salts in ILs with anions with coordinating capabilities. Ln-containing ionic liquids proved to be promising materials because, although liquids, they provide a low-phonon environment for the Ln3+ center, leading to appreciable excitation state lifetimes. Typically, liquid-state lanthanide compounds present lower emission quantum yields (Φ) than those in the solid state, due to a less rigid environment and energy loss from collisions. As such, it is not surprising that emission quantum yields for Ln@ILs are very low. The same reasoning is applicable to Ln-ILs although, surprisingly, out of the 58 Ln-ILs presented here, only one RTIL—[C6mim]3[Sm(NO3)6]—had its emission quantum yield determined, with a value of 2.73% [62]. Another important aspect to stress is that since Ln@ILs are liquids, no structural characterization was available for the majority of these compounds.
It is worth mentioning that many of the Ln-ILs were studied not only as phosphors, but also as paramagnetic liquids, opening avenues for multifunctional applications. Additionally, the combination of Ln and ILs has aroused so much interest that it led to the emergence of a new field of research, focused specifically on soft materials. In this area, new ionogels have been developed through covalently grafting—or simply dispersing—Ln complexes into silica-based materials, polymer matrices, liquid crystals, etc.
It was not intended to include ionogels in this review but, just as an example, a simple and environmentally friendly (solvent-free) preparation of ionogels via the incorporation of Ln-ILs within poly(methyl methacrylate) (PMMA) was reported by Wang et al. as early as 2013 [73]. In that work, the ILs Tb(sal)@[Carb-mim][Tf2N] and Eu(tta)@[Carb-mim] [Tf2N] were directly dissolved into MMA monomers with azodiisobutyronitrile (polymerization initiator), with stirring at 80 °C, yielding a yellowish liquid that was then cast into glass slides or glass bottles. After drying, ionogels in the form of monoliths, films, and flexible self-standing films could be obtained (Figure 14).
The accomplishments described in this review have proven Ln-ILs to be outstanding and promising optical materials. However, this field of research is still underdeveloped when compared with other fields of ionic liquid chemistry. Therefore, new studies focusing on different combinations of Ln ions and new ligands will certainly lead to more efficient luminescent molecular devices, paving the way for practical applications as varied as catalysis, biochemical analysis, energy production, and non-invasive diagnostics, such as biolabels.

Author Contributions

Conceptualization, investigation, and writing, C.C.L.P., J.M.C., B.M., and J.P.L. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the Associated Laboratory for Sustainable Chemistry-Clean Processes and Technologies—LAQV, which is financed by national funds from FCT/MEC (UID/QUI/50006/2019) and co-financed by the ERDF under the PT2020 Partnership Agreement (POCI-01-0145-FEDER–007265). This work has been supported by the Fundação para a Ciência e a Tecnologia through the contract nº IST-ID/077/2018 (B.M.) and Norma transitória DL 57/2016 Contract Program (C.C.L.P). Centro de Química Estrutural (CQE) acknowledges the financial support of the Fundação para a Ciência e Tecnologia (UIDB/00100/2020).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

Acronyms

AAAcetylacetonate
BABenzoate
betaineN,N,N-trimethylglycine
BF4Tetrafluoroborate
bmpyr1-N-butyl-1-methylpyrrolidinium
BrMAConjugate base of bromomalonaldehyde
[Carb-mim]3-(3-Carboxy-propyl)-1-methylimidazolium
[3-5-carb-mim]3-(5-Carboxy propyl) -1-methylimidazolium
Cnmim1-Alkyl-3-methylimidazolium
C2mim1-Ethyl-3-methylimidazolium
C3mim1-Propyl-3-methylimidazolium
C4mim1-Butyl-3-methylimidazolium
C4mpyN-butyl-4-methylpyridinium
C4mpyr1-Butyl-1-methylpyrrolidinium
C6mim1-Hexyl-3-methylimidazolium
C8mim1-Octyl-3-methylimidazolium
C10mim1-Decyl-3-methylimidazolium
C12mim1-Dodecyl-3-methylimidazolium
C12mpyrN-dodecyl-N-methylpyrrolidinium
DCADicyanamide
choline(2-Hydroxyethyl)trimethyl ammonium
dbm1,3-Diphenylpropane-1,3-dione
dpaPyridine-2,6-dicarboxylate
dppim1,3-Bis-{3-(diphenylphosphinyl)propyl}imidazole
hfa1,1,1,5,5,5-Hexafluoroacetylacetonate
MC1mim1,2,3-Trimethylimidazolium
MILMagnetic ionic liquid
MOEmim1-(2-Methoxyethyl)-3-methylimidazolium
NIRNear-infrared
ntaNaphthoyltrifluoroacetone
P8,8,8,1Trioctylmethylphosphonium
phen1,10-Phenantroline
PMMAPoly(methyl methacrylate)
pybox2,6-Bis[(4R)-4-phenyl-2-oxazolinyl]pyridine
RTILRoom-temperature ionic liquid
sacSaccharinate
salSalicylate
Tf2NBis-(trifluoromethanesulfonyl)amide
TfOTrifluoromethanesulfonate
tta2-Thenoyltrifluoroacetonate
TODGAN,N,N’,N’-tetra(n-octyl)diglycolamide

References

  1. Lei, Z.-G.; Chen, B.-H.; Koo, Y.-M.; MacFarlane, D.R. Introduction: Ionic Liquids. Chem. Rev. 2017, 117, 6633–6635. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  2. Prodius, D.; Smetana, V.; Steinberg, S.; Wilk-Kozubek, M.; Mudryk, Y.; Pecharsky, V.K.; Mudring, A.-V. Breaking the paradigm: Record quindecim charged magnetic ionic liquids. Mater. Horiz. 2017, 4, 217–221. [Google Scholar] [CrossRef] [Green Version]
  3. Welton, T. Room-temperature ionic liquids. Solvents for synthesis and catalysis. Chem. Rev. 1999, 99, 2071–2083. [Google Scholar] [CrossRef] [PubMed]
  4. Wilkes, J.S. A short history of ionic liquids—From molten salts to neoteric solvents. Green Chem. 2002, 4, 73–80. [Google Scholar] [CrossRef]
  5. Giernoth, R. Task-Specific Ionic Liquids. Angew. Chem. Int. Ed. 2010, 49, 2834–2839. [Google Scholar] [CrossRef]
  6. Zhao, H.; Xia, S.Q.; Ma, P.S. Use of ionic liquids as ‘green’ solvents for extractions. J. Chem. Technol. Biotechnol. 2005, 80, 1089–1096. [Google Scholar] [CrossRef]
  7. Armand, M.; Endres, F.; MacFarlane, D.R.; Ohno, H.; Scrosati, B. Ionic-liquid materials for the electrochemical challenges of the future. Nat. Mater. 2009, 8, 621–629. [Google Scholar] [CrossRef]
  8. Berthod, A.; Ruiz-Angel, M.; Carda-Broch, S. Ionic liquids in separation techniques. J. Chromatogr. A 2008, 1184, 6–18. [Google Scholar] [CrossRef]
  9. Sethurajan, M.; van Hullebusch, E.D.; Fontana, D.; Akcil, A.; Deveci, H.; Batinic, B.; Leal, J.P.; Gasche, T.A.; Kucuker, M.A.; Kuchta, K.; et al. Recent advances on hydrometallurgical recovery of critical and precious elements from end-of-life electronic wastes-a review. Crit. Rev. Env. Sci. Tech. 2019, 49, 212–275. [Google Scholar] [CrossRef] [Green Version]
  10. Sun, X.; Luo, H.; Dai, S. Mechanistic investigation of solvent extraction based on anion-functionalized ionic liquids for selective separation of rare-earth ions. Dalton Trans. 2013, 42, 8270–8275. [Google Scholar] [CrossRef]
  11. Maria, L.; Cruz, A.; Carretas, J.M.; Monteiro, B.; Galinha, C.; Gomes, S.S.; Araujo, M.F.; Paiva, I.; Marcalo, J.; Leal, J.P. Improving the selective extraction of lanthanides by using functionalised ionic liquids. Sep. Purif. Technol. 2020, 237, 116354. [Google Scholar] [CrossRef]
  12. Park, J.; Jung, Y.; Kusumah, P.; Lee, J.; Kwon, K.; Lee, C.K. Application of ionic liquids in hydrometallurgy. Int. J. Mol. Sci. 2014, 15, 15320–15343. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  13. Bermudez, M.D.; Jimenez, A.E.; Sanes, J.; Carrion, F.J. Ionic Liquids as Advanced Lubricant Fluids. Molecules 2009, 14, 2888–2908. [Google Scholar] [CrossRef]
  14. Rooney, D.; Jacquemin, J.; Gardas, R. Thermophysical Properties of Ionic Liquids. In Ionic Liquids. Topics in Current Chemistry; Kirchner, B., Ed.; Springer: Berlin, Germany, 2009; Volume 290, pp. 185–212. [Google Scholar]
  15. Mudring, A.-V. Optical Spectroscopy and Ionic Liquids. In Ionic Liquids. Topics in Current Chemistry; Kirchner, B., Ed.; Springer: Berlin, Germany, 2009; Volume 290, pp. 285–310. [Google Scholar]
  16. Tang, S.-F.; Mudring, A.-V. Highly Luminescent Ionic Liquids Based on Complex Lanthanide Saccharinates. Inorg. Chem. 2019, 58, 11569–11578. [Google Scholar] [CrossRef]
  17. Cotton, S. Lanthanide and Actinide Chemistry; John Wiley & Sons Ltd.: Hoboken, NJ, USA, 2006; pp. 1–263. [Google Scholar]
  18. Bünzli, J.C.G.; Piguet, C. Taking advantage of luminescent lanthanide ions. Chem. Soc. Rev. 2005, 34, 1048–1077. [Google Scholar] [CrossRef] [PubMed]
  19. Lei, N.; Shen, D.; Wanga, J.; Chen, X. Flexible and enhanced multicolor-emitting films co-assembled by lanthanide complexes and a polymerizable surfactant in aqueous solution. Soft Matter 2018, 14, 9143–9152. [Google Scholar] [CrossRef] [PubMed]
  20. Prodius, D.; Mudring, A.-V. Rare earth metal-containing ionic liquids. Coord. Chem. Rev. 2018, 363, 1–16. [Google Scholar] [CrossRef] [Green Version]
  21. Weissman, S.I. Intramolecular Energy Transfer the Fluorescence of Complexes of Europium. J. Chem. Phys. 1942, 10, 214–217. [Google Scholar] [CrossRef]
  22. Bünzli., J.-C.G. On the design of highly luminescent lanthanide complexes. Coord. Chem. Rev. 2015, 293, 19–47. [Google Scholar] [CrossRef]
  23. Ning, Y.; Zhu, M.; Zhang, J.-L. Near-infrared (NIR) lanthanide molecular probes for bioimaging and biosensing. Coord. Chem. Rev. 2019, 399, 213028. [Google Scholar] [CrossRef]
  24. Varaksina, E.A.; Taydakov, I.V.; Ambrozevich, S.A.; Selyukov, A.S.; Lyssenko, K.A.; Jesus, L.T.; Freire, R.O. Influence of fluorinated chain length on luminescent properties of Eu3+ β-diketonate complexes. J. Lumin. 2018, 196, 161–168. [Google Scholar] [CrossRef]
  25. Driesen, K.; Nockemann, P.; Binnemans, K. Ionic liquids as solvents for near-infrared emitting lanthanide complexes. Chem. Phys. Lett. 2004, 395, 306–310. [Google Scholar] [CrossRef]
  26. Eliseeva, S.V.; Bünzli, J.-C.G. Lanthanide luminescence for functional materials and bio-sciences. Chem. Soc. Rev. 2010, 39, 189–227. [Google Scholar] [CrossRef]
  27. Mudring, A.-V.; Tang, S. Ionic Liquids for Lanthanide and Actinide Chemistry. Eur. J. Inorg. Chem. 2010, 2569–2581. [Google Scholar] [CrossRef]
  28. Feng, J.; Zhang, H. Hybrid materials based on lanthanide organic complexes: A review. Chem. Soc. Rev. 2013, 42, 387–410. [Google Scholar] [CrossRef] [PubMed]
  29. Santos, E.; Albo, J.; Irabien, A. Magnetic ionic liquids: Synthesis, properties and applications. RSC Adv. 2014, 4, 40008–40018. [Google Scholar] [CrossRef]
  30. Cheng, K.-L.; Yuan, W.-L.; He, L.; Tang, N.; Jian, H.-M.; Zhao, Y.; Qin, S.; Tao, G.-H. Fluorescigenic Magnetofluids Based on Gadolinium, Terbium, and Dysprosium-Containing Imidazolium Salts. Inorg. Chem. 2018, 57, 6376–6390. [Google Scholar] [CrossRef]
  31. Ramos, T.J.S.; Berton, G.H.; Júnior, S.A.; Cassol, T.M. Photostable soft materials with tunable emission based on sultone functionalized ionic liquid and lanthanides ions. J. Lumin. 2019, 209, 208–216. [Google Scholar] [CrossRef]
  32. Binnemans, K. Interpretation of europium(III) spectra. Coord. Chem. Rev. 2015, 295, 1–45. [Google Scholar] [CrossRef] [Green Version]
  33. Tang, S.; Babai, A.; Mudring, A.-V. Europium-based ionic liquids as luminescent soft materials. Angew. Chem. Int. Ed. 2008, 47, 7631–7634. [Google Scholar] [CrossRef] [PubMed]
  34. Puntus, L.N.; Pekareva, I.S.; Lyssenko, K.A.; Shaplov, A.S.; Lozinskaya, E.I.; Zdvizhkov, A.T.; Buzin, M.I.; Vygodskii, Y.S. Influence of ionic liquid anion nature on the properties of Eu-containing luminescent materials. Opt. Mater. 2010, 32, 707–710. [Google Scholar] [CrossRef]
  35. Bruno, S.M.; Ferreira, R.A.S.; Paz, F.A.A.; Carlos, L.D.; Pillinger, M.; Ribeiro-Claro, P.; Gonçalves, I.S. Structural and Photoluminescence Studies of a Europium(III) Tetrakis (beta-diketonate) Complex with Tetrabutylammonium, Imidazolium, Pyridinium and Silica-Supported Imidazolium Counterions. Inorg. Chem. 2009, 48, 4882–4895. [Google Scholar] [CrossRef]
  36. Guillet, E.; Imbert, D.; Scopelliti, R.; Bünzli, J.-C.G. Tuning the emission color of europium-containing ionic liquid-crystalline phases. Chem. Mater. 2004, 16, 4063–4070. [Google Scholar] [CrossRef]
  37. Nockemann, P.; Beurer, E.; Driesen, K.; Van Deun, R.; Van Hecke, K.; Van Meervelt, L.; Binnemans, K. Photostability of a highly luminescent europium beta-diketonate complex in imidazolium ionic liquids. Chem. Commun. 2005, 34, 4354–4356. [Google Scholar] [CrossRef]
  38. Lunstroot, K.; Nockemann, P.; Van Hecke, K.; Van Meervelt, L.; Gorller-Walrand, C.; Binnemans, K.; Driesen, K. Visible and Near-Infrared Emission by Samarium(III)-Containing Ionic Liquid Mixtures. Inorg. Chem. 2009, 48, 3018–3026. [Google Scholar] [CrossRef]
  39. Babai, A.; Mudring, A.-V. Anhydrous praseodymium salts in the ionic liquid [bmpyr][Tf2N]: Structural and optical properties of [bmpyr]4[PrI6][Tf2N] and [bmyr]2[Pr(Tf2N)5]. Chem. Mater. 2005, 17, 6230–6238. [Google Scholar] [CrossRef]
  40. Arenz, S.; Babai, A.; Binnemans, K.; Driesen, K.; Giernoth, R.; Mudring, A.-V.; Nockemann, P. Intense near-infrared luminescence of anhydrous lanthanide(III) iodides in an imidazolium ionic liquid. Chem. Phys. Lett. 2005, 402, 75–79. [Google Scholar] [CrossRef]
  41. Mudring, A.-V.; Babai, A.; Arenz, S.; Giernoth, R.; Binnemans, K.; Driesen, K.; Nockemann, P. Strong luminescence of rare earth compounds in ionic liquids: Luminescent properties of lanthanide(III) iodides in the ionic liquid 1-dodecyl-3-methylimidazolium bis(trifluoromethanesulfonyl)imide. J. Alloys Compd. 2006, 418, 204–208. [Google Scholar] [CrossRef]
  42. Puntus, L.N.; Schenk, K.J.; Bünzli, J.-C.G. Intense near-infrared luminescence of a mesomorphic ionic liquid doped with lanthanide beta-diketonate ternary complexes. Eur. J. Inorg. Chem. 2005, 4739–4744. [Google Scholar] [CrossRef]
  43. Lunstroot, K.; Baeten, L.; Nockemann, P.; Martens, J.; Verlooy, P.; Ye, X.; Gorller-Walrand, C.; Binnemans, K.; Driesen, K. Luminescence of LaF3:Ln3+ Nanocrystal Dispersions in Ionic Liquids. J. Phys. Chem. C 2009, 113, 13532–13538. [Google Scholar] [CrossRef] [Green Version]
  44. Getsis, A.; Mudring, A.-V. Lanthanide Containing Ionic Liquid Crystals: EuBr2, SmBr3, TbBr3 and DyBr3 in C12mimBr. Z. Anorg. Allg. Chem. 2010, 636, 1726–1734. [Google Scholar] [CrossRef]
  45. Puntus, L.; Zhuravlev, K.; Pekareva, I.; Lyssenko, K.; Zolin, V. Peculiarities of the structure of lanthanide chloride complexes with heterocyclic diimines and the efficiency of energy transfer processes. Opt. Mater. 2008, 30, 806–809. [Google Scholar] [CrossRef]
  46. Huanrong, L.; Liu, P.; Shao, H.; Wang, Y.; Zheng, Y.; Sunc, Z.; Chen, Y. Green synthesis of luminescent soft materials derived from task-specific ionic liquid for solubilizing lanthanide oxides and organic ligand. J. Mater. Chem. 2009, 19, 5533–5540. [Google Scholar]
  47. Hopkins, T.; Goldey, M. Tb3+ and Eu3+ luminescence in imidazolium ionic liquids. J. Alloys Compd. 2009, 488, 615–618. [Google Scholar] [CrossRef]
  48. Brandner, A.; Kitahara, T.; Beare, N.; Lin, C.; Berry, M.; May, P. Luminescence Properties and Quenching Mechanisms of Ln(Tf2N)3 Complexes in the Ionic Liquid bmpyr Tf2N. Inorg. Chem. 2011, 50, 6509–6520. [Google Scholar] [CrossRef]
  49. Devi, V.; Maji, S.; Viswanathan, K. Novel room temperature ionic liquid for fluorescence enhancement of Eu3+ and Tb3+. J. Lumin. 2011, 131, 739–748. [Google Scholar] [CrossRef]
  50. Getsis, A.; Mudring, A.-V. Switchable Green and White Luminescence in Terbium-Based Ionic Liquid Crystals. Eur. J. Inorg. Chem. 2011, 3207–3213. [Google Scholar] [CrossRef]
  51. Li, H.-R.; Li, D.; Wang, Y.-G.; Ru, Q.-R. A Series of Carboxylic-Functionalized Ionic Liquids and their Solubility for Lanthanide Oxides. Chem. Asian J. 2011, 6, 1443–1449. [Google Scholar] [CrossRef] [PubMed]
  52. Bortoluzzi, M.; Battistel, D.; Roppa, S.; Daniele, S.; Perosa, A.; Enrichi, F. Yttrium and lanthanide complexes of beta-dialdehydes: Synthesis, characterization, luminescence and electrochemistry of coordination compounds with the conjugate base of bromomalonaldehyde. Dalton Trans. 2014, 43, 9303–9312. [Google Scholar] [CrossRef]
  53. Malba, C.M.; Enrichi, F.; Facchin, M.; Demitri, N.; Plaisier, J.R.; Natile, M.M.; Selva, M.; Riello, P.; Perosa, A.; Benedetti, A. Phosphonium-based tetrakis dibenzoylmethane Eu(iii) and Sm(iii) complexes: Synthesis, crystal structure and photoluminescence properties in a weakly coordinating phosphonium ionic liquid. RSC Adv. 2015, 5, 60898–60907. [Google Scholar] [CrossRef]
  54. Zhang, X.-Y.; Wang, T.-R.; Qin, X.-T.; Zhang, Z.-Y.; Sun, Y.-Y.; Liang, H.-Z.; Li, H.-R. Large-area flexible, transparent, and highly luminescent films containing lanthanide (III) complex-doped ionic liquids for efficiency enhancement of silicon-based heterojunction solar cell. Prog. Photovolt. 2017, 25, 1015–1021. [Google Scholar] [CrossRef]
  55. Tang, S.-F.; Smetana, V.; Mishra, M.K.; Kelley, S.P.; Renier, O.; Rogers, R.D.; Mudring, A.-V. Forcing Dicyanamide Coordination to f-Elements by Dissolution in Dicyanamide-Based Ionic Liquids. Inorg. Chem. 2020, 59, 7227–7237. [Google Scholar] [CrossRef] [PubMed]
  56. Mallick, B.; Balke, B.; Felser, C.; Mudring, A.-V. Dysprosium Room-Temperature Ionic Liquids with Strong Luminescence and Response to Magnetic Fields. Angew. Chem. Int. Ed. 2008, 47, 7635–7638. [Google Scholar] [CrossRef]
  57. Jensen, M.P.; Neuefeind, J.; Beitz, J.V.; Skanthakumar, S.; Soderholm, L. Mechanisms of Metal Ion Transfer into Room-Temperature Ionic Liquids: The Role of Anion Exchange. J. Am. Chem. Soc. 2003, 125, 15466–15473. [Google Scholar] [CrossRef]
  58. Nockemann, P.; Thijs, B.; Postelmans, N.; Van Hecke, K.; Van Meervelt, L.; Binnemans, K. Anionic Rare-Earth Thiocyanate Complexes as Building Blocks for Low-Melting Metal-Containing Ionic Liquids. J. Am. Chem. Soc. 2006, 128, 13658–13659. [Google Scholar] [CrossRef]
  59. Ohaion, T.; Kalisky, Y.; Ben-Eliyahu, Y.; Becker, J.Y.; Bettelheim, A. Spectral and Electrochemical Properties of Lanthanide Thiocyanate Complexes as Ionic Liquid Components. Eur. J. Inorg. Chem. 2013, 3477–3482. [Google Scholar] [CrossRef]
  60. Getsis, A.; Balke, B.; Felser, C.; Mudring, A.V. Dysprosium-Based Ionic Liquid Crystals: Thermal, Structural, Photo- and Magnetophysical Properties. Cryst. Growth Des. 2009, 9, 4429–4437. [Google Scholar] [CrossRef]
  61. Li, M.; Ganea, G.M.; Lu, C.; De Rooy, S.L.; El-Zahab, B.; Fernand, V.E.; Jin, R.; Aggarwal, S.; Warner, I.M. Lipophilic phosphonium–lanthanide compounds with magnetic, luminescent, and tumor targeting properties. J. Inorg. Biochem. 2012, 107, 40–46. [Google Scholar] [CrossRef]
  62. Tang, N.; Zhao, Y.; He, L.; Yuan, W.-L.; Tao, G.-H. Long-lived luminescent soft materials of hexanitratosamarate(III) complexes with orange visible emission. Dalton Trans. 2015, 44, 8816–8823. [Google Scholar] [CrossRef] [PubMed]
  63. Han, Y.-L.; Lin, C.-K.; Meng, Q.-G.; Dai, F.-R.; Sykes, A.G.; Berry, M.T.; May, P.S. (BMI)3LnCl6 Crystals as Models for the Coordination Environment of LnCl3 (Ln = Sm, Eu, Dy, Er, Yb) in 1-Butyl-3-methylimidazolium Chloride Ionic-Liquid Solution. Inorg. Chem. 2014, 53, 5494–5501. [Google Scholar] [CrossRef]
  64. Fan, B.; Wei, J.; Ma, X.; Bu, X.; Xing, N.; Pan, Y.; Zheng, L.; Guan, W. Synthesis of Lanthanide-Based Room Temperature Ionic Liquids with Strong Luminescence and Selective Sensing of Fe(III) over Mixed Metal Ions. Ind. Eng. Chem. Res. 2016, 55, 2267–2271. [Google Scholar] [CrossRef]
  65. Pohako-Esko, K.; Wehner, T.; Schulz, P.S.; Heinemann, F.W.; Müller-Buschbaum, K.; Wasserscheid, P. Synthesis and Properties of Organic Hexahalocerate(III) Salts. Eur. J. Inorg. Chem. 2016, 1333–1339. [Google Scholar] [CrossRef]
  66. Alvarez-Vicente, J.; Dandil, S.; Banerjee, D.; Gunaratne, H.Q.N.; Gray, S.; Felton, S.; Srinivasan, G.; Kaczmarek, A.M.; Deun, R.V.; Nockemann, P. Easily Accessible Rare-Earth-Containing Phosphonium Room-Temperature Ionic Liquids: EXAFS, Luminescence, and Magnetic Properties. J. Phys. Chem. B 2016, 120, 5301–5311. [Google Scholar] [CrossRef] [Green Version]
  67. Nockemann, P.; McCourt, E.; Esien, K.; Felton, S.; Zhenyu, L. Designing Dimeric Lanthanide(III)-Containing Ionic liquids. Angew. Chem. Int. Ed. 2018, 2021. [Google Scholar] [CrossRef] [Green Version]
  68. Pierson, S.A.; Nacham, O.; Clark, K.D.; Nan, H.; Mudryk, Y.; Anderson, J.L. Synthesis and characterization of low viscosity hexafluoroacetylacetonate-based hydrophobic magnetic ionic liquids. New J. Chem. 2017, 41, 5498–5505. [Google Scholar] [CrossRef]
  69. Monteiro, B.; Outis, M.; Cruz, H.; Leal, J.P.; Laia, C.A.T.; Pereira, C.C.L. A thermochromic europium (III) room temperature ionic liquid with thermally activated anion–cation interactions. Chem. Commun. 2017, 53, 850–853. [Google Scholar] [CrossRef]
  70. Leal, J.P.; Otis, M.; Casimiro, M.H.; Ferreira, L.M.; Fernandes, F.; Monteiro, B.; Laia, C.A.T.; Pereira, C.C.L. A case of self-organization in highly emissive Eu(III) ionic liquids. Eur. J. Inorg. Chem. 2017, 3429–3434. [Google Scholar] [CrossRef]
  71. Zheng, L.; Yang, L.-L.; Xing, N.-N.; Pan, Y.; Ji, H.-X.; Wei, J.; Guan, W. Highly selective detection of nitrotoluene based on novel lanthanide-containing ionic liquids. RSC Adv. 2017, 7, 35814–35818. [Google Scholar] [CrossRef] [Green Version]
  72. Wu, K.; Shen, X. Designing a new type of magnetic ionic liquid: A strategy to improve the magnetic susceptibility. New J. Chem. 2019, 43, 15857–15860. [Google Scholar] [CrossRef]
  73. Wang, H.; Wang, Y.; Zhang, L.; Li, H. Transparent and luminescent ionogels based on lanthanide-containing ionic liquids and poly(methylmethacrylate) prepared through an environmentally friendly method. RSC Adv. 2013, 3, 8535–8540. [Google Scholar] [CrossRef]
Figure 1. Lanthanide family, excluding Promethium (Pm with atomic number 61 does not exist in nature). Eu (not covered in this paper) is in blue; lighter elements in brown and heavier elements in green.
Figure 1. Lanthanide family, excluding Promethium (Pm with atomic number 61 does not exist in nature). Eu (not covered in this paper) is in blue; lighter elements in brown and heavier elements in green.
Molecules 26 04834 g001
Figure 2. (a) Schematic representation of the antenna effect, and (b) a Jablonski diagram of the antenna effect (solid arrows) together with competing energy transitions (dashed arrows). Abs: absorbance; IC: internal conversion; Fluor: fluorescence; ISC: intersystem crossing; Phos: phosphorescence; ET: energy transfer; EbT: energy back-transfer; Lum: luminescence.
Figure 2. (a) Schematic representation of the antenna effect, and (b) a Jablonski diagram of the antenna effect (solid arrows) together with competing energy transitions (dashed arrows). Abs: absorbance; IC: internal conversion; Fluor: fluorescence; ISC: intersystem crossing; Phos: phosphorescence; ET: energy transfer; EbT: energy back-transfer; Lum: luminescence.
Molecules 26 04834 g002
Figure 3. Energy diagram of the NIR luminescent states of Nd3+, Er3+, and Yb3+, and C–H, N–H, and O–H overtones.
Figure 3. Energy diagram of the NIR luminescent states of Nd3+, Er3+, and Yb3+, and C–H, N–H, and O–H overtones.
Molecules 26 04834 g003
Figure 4. Luminescence of [C6mim][Sm(nta)4] (nta = 2-naphthoyltrifluoroacetonate) dissolved in the ionic liquid [C6mim][Tf2N] under UV light (λ = 365 nm) irradiation. Reproduced with permission from [38].
Figure 4. Luminescence of [C6mim][Sm(nta)4] (nta = 2-naphthoyltrifluoroacetonate) dissolved in the ionic liquid [C6mim][Tf2N] under UV light (λ = 365 nm) irradiation. Reproduced with permission from [38].
Molecules 26 04834 g004
Figure 5. Photographs of the flexible, transparent, luminescent PMMA films (17 × 17 cm2) under daylight and a 254-nm UV lamp. Adapted with permission from [54].
Figure 5. Photographs of the flexible, transparent, luminescent PMMA films (17 × 17 cm2) under daylight and a 254-nm UV lamp. Adapted with permission from [54].
Molecules 26 04834 g005
Figure 6. CIE 1931 chromaticity diagram for Ln compounds (red, (Eu3+); green, (Tb3+); blue, (Dy3+); circle: at RT, and square: at LT (77 K)). Reproduced with permission from [55].
Figure 6. CIE 1931 chromaticity diagram for Ln compounds (red, (Eu3+); green, (Tb3+); blue, (Dy3+); circle: at RT, and square: at LT (77 K)). Reproduced with permission from [55].
Molecules 26 04834 g006
Figure 7. Visualization of live cancer cells in the presence of [Ph4P]5[Dy(SCN)8] by confocal laser microscopy: (a) localization of the nucleus (control test showing cell nucleus without using Dy-IL); (b) localization of the Dy-IL; (c) (a) and (b) overlapped; (d) demonstration of how Dy-IL was taken by the cells up (λex = 458 nm). Reproduced with permission from [61].
Figure 7. Visualization of live cancer cells in the presence of [Ph4P]5[Dy(SCN)8] by confocal laser microscopy: (a) localization of the nucleus (control test showing cell nucleus without using Dy-IL); (b) localization of the Dy-IL; (c) (a) and (b) overlapped; (d) demonstration of how Dy-IL was taken by the cells up (λex = 458 nm). Reproduced with permission from [61].
Molecules 26 04834 g007
Figure 8. Normalized emission spectra of (a) 1% SmCl3 in C4mimCl and [C4mim]3[SmCl6] crystals; (b) 1% DyCl3 in C4mimCl and [C4mim]3[GdCl6]:[5% Dy3+] crystals; (c) 1% ErCl3 in [C4mim][Cl] and [C4mim]3[ErCl6] crystals; and (d) 1% YbCl3 in [C4mim][Cl] and [C4mim]3[YbCl6] crystals. Solution and crystal spectra were measured at 338 K and 278 K, respectively. Reproduced with permission from [63].
Figure 8. Normalized emission spectra of (a) 1% SmCl3 in C4mimCl and [C4mim]3[SmCl6] crystals; (b) 1% DyCl3 in C4mimCl and [C4mim]3[GdCl6]:[5% Dy3+] crystals; (c) 1% ErCl3 in [C4mim][Cl] and [C4mim]3[ErCl6] crystals; and (d) 1% YbCl3 in [C4mim][Cl] and [C4mim]3[YbCl6] crystals. Solution and crystal spectra were measured at 338 K and 278 K, respectively. Reproduced with permission from [63].
Molecules 26 04834 g008
Figure 9. Comparison of the photoluminescence intensity of (a) [C4mim][Dy(NO3)4] and (b) [C4mim][Dy(NO3)4] mixed with Mn+ (Ca(II), Al(III), Zn(II), Cu(II), Pb(II), Hg(II), Cd(II), Co(II), Fe(II), Ni(II), and Cr(III)), in the absence and presence of Fe(III). Insets are the corresponding luminescence images under UV light irradiation at 365 nm. Reproduced with permission from [64].
Figure 9. Comparison of the photoluminescence intensity of (a) [C4mim][Dy(NO3)4] and (b) [C4mim][Dy(NO3)4] mixed with Mn+ (Ca(II), Al(III), Zn(II), Cu(II), Pb(II), Hg(II), Cd(II), Co(II), Fe(II), Ni(II), and Cr(III)), in the absence and presence of Fe(III). Insets are the corresponding luminescence images under UV light irradiation at 365 nm. Reproduced with permission from [64].
Molecules 26 04834 g009
Figure 10. Structure of the [Er5(C2H5-C3H3N2-CH2COO)16(H2O)8]15+ cation. Reproduced with permission from [2].
Figure 10. Structure of the [Er5(C2H5-C3H3N2-CH2COO)16(H2O)8]15+ cation. Reproduced with permission from [2].
Molecules 26 04834 g010
Figure 11. Images of a U-tube filled with Tb-ILs (a–d) and Dy-IL (e–h), with (b,d,f,h) and without (a,c,e,g) a NdFeB magnet on the side under the radiation of bright light (a,b,e,f) and 365 nm UV light (c,d,g,h). Reproduced with permission from [30].
Figure 11. Images of a U-tube filled with Tb-ILs (a–d) and Dy-IL (e–h), with (b,d,f,h) and without (a,c,e,g) a NdFeB magnet on the side under the radiation of bright light (a,b,e,f) and 365 nm UV light (c,d,g,h). Reproduced with permission from [30].
Molecules 26 04834 g011
Figure 12. (AD) Photographs of ionic liquid 1,4-methylimidazolylbutane sulfonate bromide (white emission) and Ln-ILs: (E,F) Gd(IL-1)3(H2O)3, (G,H) Eu(IL-1)3(H2O)3 (red emission), and (I,J) Gd(IL-1)3(H2O)3 (green emission) under UV light (365 nm) and daylight. Reproduced with permission from [31].
Figure 12. (AD) Photographs of ionic liquid 1,4-methylimidazolylbutane sulfonate bromide (white emission) and Ln-ILs: (E,F) Gd(IL-1)3(H2O)3, (G,H) Eu(IL-1)3(H2O)3 (red emission), and (I,J) Gd(IL-1)3(H2O)3 (green emission) under UV light (365 nm) and daylight. Reproduced with permission from [31].
Molecules 26 04834 g012
Figure 13. Emission spectra of (a) Tb(TODGA)3Cl3, (b) NH4[Tb(hfa)4], and (c) [Tb(TODGA)3][Tb(hfa)4]3 in 1-mM acetonitrile solutions. Reproduced with permission from [72].
Figure 13. Emission spectra of (a) Tb(TODGA)3Cl3, (b) NH4[Tb(hfa)4], and (c) [Tb(TODGA)3][Tb(hfa)4]3 in 1-mM acetonitrile solutions. Reproduced with permission from [72].
Molecules 26 04834 g013
Figure 14. Digital photos of the Ln-ILs-PMMA: (ac) Eu(tta)–[Carb-mim][Tf2N]@PMMA under daylight (right) and UV light (left); (de) Tb(sal)–[Carb-mim][Tf2N]@PMMA under daylight (right) and UV light (left). The scale bar is 1.0 cm. Reproduced with permission from [73].
Figure 14. Digital photos of the Ln-ILs-PMMA: (ac) Eu(tta)–[Carb-mim][Tf2N]@PMMA under daylight (right) and UV light (left); (de) Tb(sal)–[Carb-mim][Tf2N]@PMMA under daylight (right) and UV light (left). The scale bar is 1.0 cm. Reproduced with permission from [73].
Molecules 26 04834 g014
Table 1. List of Ln@ILs and their lifetimes and physical states/transitions.
Table 1. List of Ln@ILs and their lifetimes and physical states/transitions.
Ln@ILsLifetimes
(Excited Level,
Decay Lifetime)
Absolute
Quantum Yield (%)
Ref.
Pr(Tf2N)3@[bmpyr][Tf2N]3P0, 60.0 ns
1D2, 1.8 µs
0.0013148
PrI3@[bmpyr][Tf2N]3P239
Pr(Tf2N)3@[bmpyr][Tf2N] 3P239
Nd(Tf2N)3@[bmpyr][Tf2N]4F3/2; 25.3 µs48
LaF3:5%Nd3+:betaine@[C4mpyr][Tf2N]4F3/243
LaF3:5%Nd3+:betaine@[C4mpyr][TfO]4F3/243
LaF3:5%Nd3+:betaine@[C4mim][Tf2N]4F3/243
Nd(tta)@ [3-5-carb-mim][Br]4F3/246
Nd(tta)@ [3-5-carb-mim][Br]4F3/246
[Nd(tta)3(phen)]@ [C12mim][Cl]4F3/2; 1.52 µs42
NdI3@[C12mim][Tf2N]4F3/2; 15.3 µs1.0540
Nd(TfO)3@[EMIM][TfO] 4G5/2; 376 ns25
Nd(TOS)3@[EMIM][TOS]4G5/2; 800 ns25
NdBr3@[HMIM]Br]4G5/2; 1510 ns25
Nd(PBS)3(Phen)@[HMIM](PBS]4G5/2; 260 ns25
Nd(NTA)4@[HMIM]Br]4G5/2; 1400 ns25
Sm(Tf2N)3@[bmpyr][Tf2N]4G5/2; 259 µs0.06048
SmBr3@[C12mim]Br4G5/2; -9°C 105°C44
[C6mim][Sm(tta)4]@[C6mim][Tf2N]4G5/2; 78 µs1.6038
[C6mim][Sm(nta)4]@[C6mim][Tf2N]4G5/2; 66 µs1.3738
[C6mim][Sm(hfa)4]@[C6mim][Tf2N]4G5/2; 72 µs1.0638
[choline]3[Sm(dpa)3]@[C6mim][Tf2N]4G5/2; 61 µs38
[P8,8,8,1][Sm(dbm)4]@ [P8,8,8,1][Tf2N]4G5/2, 19.1 μs0.653
TbCl3@[P8,8,8,1][BrMA]5D452
Tb4O7@[Carb-Cnminm]Br5D451
TbBr3@[C12mim]Br4D4; 3.4 ms44
TbCl3@BMIBr5D447
TbCl3(Phen)2(H2O)3@[C12mim][Cl]5D445
TbI3@[C12mim][Tf2N]5D441
Tb(pybox)3@[dppim] [bis(trifluoromethylsulfonyl)imide]5D4, 1.22 ms97.2254
[C2mim][Tb(DCA)4(H2O)4]
@[C2mim][DCA]
5D4, 0.60 ms (RT) a,
0.71 ms (LT) b
55
Dy(Tf2N)3@[bmpyr][Tf2N]4F9/2; 244 µs0.12248
DyBr3@[C12mim]Br4F9/2; 53 µs44
DyI3@[C12mim][Tf2N]4F9/2; 63 µs41
[C2mim][Dy(DCA)4(H2O)4]
@[C2mim][DCA]
4F9/2, 11.9 μs (RT)55
Er(Tf2N)3@[bmpyr][Tf2N];4S3/2, 0.15 µs
4I13/2, 70.6 µs
48
Er(tta)@ [3-5-carb-mim] [Br]4I13/246
Er(tta)@ [3-5-carb-mim] [Br]4I13/246
Er(tta)3(phen)]@[C12mim][Cl]4I13/2; 1.95 µs42
ErI3@[C12mim][Tf2N]4I13/2; 10.4 µs40
Tm(Tf2N)3@[bmpyr][Tf2N]1G4, 12.0 µs
1D2, 6.3 µs
0.20948
[Yb(tta)3(phen)]@ [C12mim][Cl]2F5/2; 12.4 µs2.142
a RT: room temperature; b LT: low temperature
Table 2. List of Ln-ILs and their published lifetimes and physical state/transitions.
Table 2. List of Ln-ILs and their published lifetimes and physical state/transitions.
Ln-ILsLifetimes
(Excited Level, Decay Lifetime)
Physical
State/Transition
(°C)
Ref.
[C10mim][CeBr6]m.p. A: 8565
[C4mim]3[CeBr3Cl3]m.p.: 10065
[C4mim][Nd(tta)4]RTIL57
[P66614]3[NdCl6]4F3/2, 2.470 μsm.p.: –45.569
[P4448]3[NdCl6]4F3/2, 2.470 μsm.p.: 46.4
Tg: –25.5
69
[P4444]3[NdCl6]4F3/2, 1.575 μsm.p.: 65.369
[C2mim]x–3[Sm(NCS)x(H2O)y]x=6, 4G5/2, ~23 μsRTIL59
x=7, 4G5/2, ~45 μs
x=8, 4G5/2, ~75 μs
[Cnmim]3[Sm(NO3)6]n= 2, 4, 8, 4G5/2,
[114.4–130.3] μs
n=6, 4G5/2, 114.8/3.75α μs
n=2, m.p.: 82, Tg: B –49
n=4, RTIL, Tg: –38
n=6, RTIL, Tg: ~41
n=8, RTIL, Tg: –45
62
[C4mim]3[SmCl6]4G5/2, 54 μs βφ63
[C4mim][Sm(NO3)4]RTIL64
[P44414]3[SmCl6]4G5/2, 33 μsm.p.: –44.969
[P4448]3[SmCl6]4G5/2, 57 μsm.p.: 47.3
Tg: –23.1
69
[P4444]3[SmCl6]4G5/2, 26 μsm.p.: 58.869
[RCnmim]2[Tb(NO3)5]R=CH3, n=1; 5D4, 1.279 ms#
R=H, n=2; 5D4, 1.294 ms
R=H, n=4; 5D4, 1.270 ms
R=H, n=6; 5D4, 1.262 ms
R=H, n=8; 5D4, 1.283 ms
m.p.: 101
RTIL, Tg: –47
RTIL, Tg: –52
RTIL, Tg: –52
RTIL, Tg: –50
30
Tb(IL-1)3(H2O)3λexc=370 nm, 5D4, 0.7712 ms
λexc=350 nm, 5D4, 0.6988 ms
RTIL (Newtonian fluid)31
[C12mim]3[TbBr6]5D4, 3.7 msS-LCC (heating): –6
LC-SD (cooling): –13
LC-LISOE (heating): 100.7
LISO-LCF (cooling): 99.4
51
6[C12mim]Br•[C12mim]3[TbBr6]5D4, 3.3 msS-LC (heating): –3.5
LC-S (cooling): –18
LC-LISO (heating): 101.5
LISO-LC (cooling): 100.9
51
[C12mpyr]3[TbBr6]5D4, 4.0 ms LC-LISO (heating): ~165
LISO-LC (cooling): ~135
51
6[C12mpyr]Br•[C12mpyr]3[TbBr6]5D4, 4.4 msS-LC (heating): 54.2
LC-S (cooling): ~30
LC-LCG (heating): ~81, ~100
LC-LISO (heating): 119
LC-LC (cooling): ~119, ~78 LISO-LC (cooling): ~75
51
[P66614]3[TbCl6]5D4, 0.692 msm.p.: –49.169
[P4448]3[TbCl6]5D4, 0.471 msTg: –25.169
[P4444]3[TbCl6]5D4, 0.416 msm.p.: 51.669
[Tb(TODGA)3][Tb(hfa)4]3RTIL, Tg: –31.0572
[RCnmim]2[Dy(NO3)5]R=CH3, n=1; 4F9/2, 54.75 μs#
R=H, n=2; 4F9/2, 59.27 μs
R=H, n=4; 4F9/2, 58.17 μs
R=H, n=6; 4F9/2, 58.21 μs
R=H, n=8; 4F9/2, 58.56 μs
m.p.: 109
m.p.: 20, Tg: –53
m.p.: 20, Tg: –55
RTIL, Tg: –55
RTIL, Tg: –55
30
[C6mim]5–x[Dy(SCN)8–x(H2O)x]x=2, 4F9/2, 23.8 μsRTIL56
x=1, 4F9/2, 40.34 μs
x=0, 4F9/2, 48.4 μs
[C12mim]3[DyBr6]room temp., 4F9/2, 47 μsLC-LC (heating): 27.8, 49.9, 87.3, 112.3, 115.5.
LC-LC (cooling): 25.7, 83.6, 112.5, 113.5.
Tg: ~20
60
70 °C, 4F9/2, 46 μs
[Ph4P]5[Dy(SCN)8]m.p.: ~3561
[Ph3PBnOEt]5[Dy(SCN)8]m.p.: ~4061
[Ph3PBnNO2]5[Dy(SCN)8]m.p.: ~4561
[Ph3PBn]5[Dy(SCN)8]m.p.: ~4061
[C4mim]3[DyCl6]4F9/2, 58 μs βφ63
[C4mim][Dy(NO3)4]RTIL64
[P66614]3[DyCl6]m.p.: –47.769
[P4448]3[DyCl6]4F9/2, 56 μsm.p.: 3.3
Tg: –31.4
69
[P4444]3[DyCl6]4F9/2, 55 μsm.p.: 42.969
[MOEmim][Dy(NO3)4]RTIL71
[Ho5(C2H5-C3H3N2-CH2COO)16(H2O)8](Tf2N)155F5, 0.8 μsmetastable RTIL
TgΔ
2
[C4mim]3[ErCl6]4I13/2, 2.5 μs βφ66
[P66614]3[ErCl6]4I13/2, 2.725 μsm.p.: –40.469
[P4448]3[ErCl6]4I13/2, 2.736 μsm.p.: –48
Tg: –67.9
69
[P4444]3[ErCl6]4I13/2, 2.066 μsm.p.: 51.369
[Er5(C2H5-C3H3N2-CH2COO)16
(H2O)8](Tf2N)15
4I13/2, 0.6 μsØm.p.: 74.6
(metastable RTIL)
TgΔ
2
[C4mim]3[YbCl6]2F5/2, 19.7 μs βφ66
A m.p.: melting point (measured on heating); B Tg: glass transition; C S–LC: solid–liquid crystal transition; D LC–S: liquid crystal–solid transition; E LC–LISO: liquid crystal–isotropic liquid transition; F LISO–LC: isotropic liquid–liquid crystal transition; G LC–LC, liquid crystal–liquid crystal transition; α the lower value was measured after addition of water; β average lifetime for the solution and crystal phases; φ studies were made in crystal form and in C4mimCl solutions, but not in pure [C4mim]3[LnCl6] form; TgΔ several glass transitions were recorded, with the temperature of the transitions gradually changing with consecutive cycles; Ø, value of lifetime measured both in liquid and solid states; # this compound is not an IL, but a low-melting-temperature salt added here for purposes of comparison.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Pereira, C.C.L.; Carretas, J.M.; Monteiro, B.; Leal, J.P. Luminescent Ln-Ionic Liquids beyond Europium. Molecules 2021, 26, 4834. https://doi.org/10.3390/molecules26164834

AMA Style

Pereira CCL, Carretas JM, Monteiro B, Leal JP. Luminescent Ln-Ionic Liquids beyond Europium. Molecules. 2021; 26(16):4834. https://doi.org/10.3390/molecules26164834

Chicago/Turabian Style

Pereira, Cláudia C. L., José M. Carretas, Bernardo Monteiro, and João P. Leal. 2021. "Luminescent Ln-Ionic Liquids beyond Europium" Molecules 26, no. 16: 4834. https://doi.org/10.3390/molecules26164834

Article Metrics

Back to TopTop