Talita Jordanna de
Souza Ramos
a,
Guilherme Henrique
Berton
b,
Tania Maria
Cassol
b and
Severino Alves
Júnior
*a
aLaboratório de Terras Raras, Departamento de Química, Centro de Ciências Exatas e da Natureza (CCEN), Universidade Federal de Pernambuco, Recife – PE, 50740-560, Brazil. E-mail: salvesjr@ufpe.br
bGrupo de Líquidos Iônicos e Metais, Departamento de Química e Biologia, Universidade Tecnológica Federal do Paraná, São Francisco Beltrão – PR, 85601-970, Brazil
First published on 18th May 2018
Herein, we describe for the first time room temperature ionic liquids (RTILs) and imidazole-based cations with appended carboxylic acids as terminals, which are directly derived from the anhydrides. All structures were designed to meet the following criteria: (1) easy preparation (up-scaling into kg scale possible); (2) melting point of <25 °C, aligned with a thermal stability of up to 100 °C; (3) chemical stability in water and common organic solvents; (4) relatively high viscosity; and (5) featured luminescence. One of the central questions in our work is how ligands can enforce a specific structure and how this enforcement may lead to several properties. Luminescent, flexible, stable, and transparent ionogels were obtained from ILs in coordination with lanthanide ions (Eu or Gd). Ionogels combine the properties of ionic liquids and show outstanding luminescence properties with decay times >1 ms, high quantum efficiency (47%), featured quantum yield (27%), and a remarkable sensitization efficiency to ILs of up to 83%, which suggests a synergistic coordination with lanthanide ions. We show that the materials obtained are potentially applicable for the construction of light-emitting electrochemical cells, owing to the properties of their basic components, including ionic liquids (with high electrical conductivity and a similar structure to liquid crystals) and lanthanide ions (which have unprecedented photophysical properties).
The introduction of rare-earth compounds into ionic liquids dates from 200611 and has aroused so much interest that there has been an emergence of a wide field of research focused specifically on soft materials.12,13 This field of research was motivated by the potential of ILs to minimize vibration-induced deactivation processes to design a rigid metal-ion environment, free of high-energy vibrations and protecting the Ln3+ ion from solvent interaction. Soft materials favourably combine the properties of ionic liquids with the unique optical properties of rare-earth compounds, such as a sharp emission band, long decay time in the excited state, and a tunable emission from the UV to the infrared spectral region. Even with the highlighted photophysical features,10 luminophores formed with several types of oxo-ligands14–16 coordinated to lanthanide ions do not show photostability. It is well known in the literature that carboxylate termini provide several coordination modes that favour the formation of three-dimensional structures.17 The primary goal of our work was to develop a new family of ionic liquids and apply them to the development of photostable structures.
All technologies under development are based on solid state electroluminescent materials and belong to the general area of solid-state lighting (SSL). The main technologies being developed in SSL are light-emitting diodes (LEDs), organic light-emitting diodes (OLEDs), and light-emitting electrochemical cells (LECs). Light-emitting electrochemical cells (LECs) have emerged as an alternative option.18 Generally, LECs have several advantages over OLEDs, such as a single-layer configuration, balanced carrier injection, low voltage operation and allowing the use of air-stable electrodes.19 The luminescent materials in LECs are either luminescent polymers together with ionic salts, or ionic species.18 The desired properties of materials for LECs are high electrical conductivity and visible light emission. Furthermore, higher electroluminescence efficiencies are expected due to the well-known phosphorescent nature of the metal complexes.20 Owing to these advantages, LECs based on metal complexes of Ir3+, Cu3+ and Ru3+ have attracted much attention in recent years.21–23 Even with the unprecedented photophysical properties of lanthanide–IL complexes, the synthesis of materials potentially applicable for the construction of light-emitting electrochemical cells has not been reported.
This report is a first example of four ionic liquids (ILs) formulated from anhydrides, imidazole-based cations with appended carboxylic acids as terminals. We examine the potential of RTILs as building blocks for coordination chemistry with Ln(III). We focus on the structure–property relationship for luminescent ionogels from experimental results and a computational study of the photoluminescence. Spectroscopic properties, such as intensity parameters Ωλ (λ = 2, 4 and 6), energy transfer (WET) and back transfer (WBT) rates, radiative (Arad) and nonradiative (Anrad) decay rates, and the quantum efficiency (η), of Eu(III) compounds were computationally modelled using the electronic and spectroscopic semi-empirical approach24 widely used in photoluminescence studies.25–28
Based on the theories developed by Malta,29,30 we calculated the 4f–4f intensity parameters and emission quantum efficiency (η). The emission quantum efficiency was determined using eqn (1), where Arad is the radiative decay rate, given by the sum of the spontaneous emission coefficients (AJs) . The total radiative decay rate (Atotal) is given by Atotal = τ−1, where τ is the lifetime for radiative decay.31 Finally, the non-radiative decay rate (Anrad) is given by Anrad = Atotal − Arad.
![]() | (1) |
The 1H NMR spectra (Fig. S4, ESI†) show behaviour qualitatively similar to the 13C NMR spectra (Fig. S5, ESI†), and both are consistent with the COSY/HSQC characterization results (Fig. S6–S9, ESI†), suggesting the formation of the proposed structures of the ILs. The NMR spectra are shown in detail in the ESI† (Tables S1 and S2). FTIR spectroscopy interpretation was performed according to Table S3 (ESI†) that includes the intensity and attribution of each stretch. As expected, functionalized IL carboxyl groups can be identified by the absorption bands at 1560 and 1484 cm−1, as shown in Fig. S10 (ESI†), which are assigned to the asymmetric stretching and symmetric stretching of the carboxylate group, respectively.32 FTIR spectra of MC1–4 showed bands at ∼1586 and 1560 cm−1, which are assigned to the imidazole ring.33 Vibrational bands in the far-IV spectra (<200 cm−1) characterize the cation–anion interaction between the imidazolium cation and bromide ions.34,35 Upon addition of Eu/Gd salts into the carboxyl-functionalized IL, the shift of absorption bands corresponding to the carboxylate moieties can be observed due to the coordination with Ln3+ ions (Fig. S11–S14, ESI†). The coordination of Ln3+ ions with ILs can be further confirmed by the luminescence data, which will be discussed later. A detailed discussion of the FTIR spectroscopy results for ligands and complexes is presented in the ESI.†
We describe a complete vibrational characterization via Raman spectroscopy, which is not commonly used for ionic liquids due to the fluorescence effect. Raman is more interesting than FTIR because the imidazolium-based ILs are highly ordered H-bonded polymeric liquids.36 However, we obtained very sharp spectra, as shown in Fig. 2. We first focus on the region between 70–300 cm−1, which is dominated by C–HA and/or N–HA (A = anion) interactions.37 As can be clearly seen in Fig. 2, the maximum intensity of the measured spectrum occurs at ∼111 cm−1 for MC3. Hydrogen bonds to the cation can be formed via N–H in the imidazolium cation and aromatic-H, explaining why the signal for MC2 is much less intense and shifted to 115 cm−1 by steric hindrance. When the structure is incapable of strong hydrogen bonding via aromatic-H, the band is shifted to 104 cm−1 (in MC1) and 100 cm−1 (in MC4), and the intensity of the peak falls. The band shift to lower wavenumbers means a reduced interaction, suggesting that the strength of interaction between cation and anion is in an escalating order such that MC2 > MC3 > MC1 > MC4. Changes in the alkyl chain length cause changes in the interactive forces, which result in a decreased electrostatic attraction between the cation and anion.38 Strong signals observed in the region at the high-frequency band (∼128 cm−1) were enlarged due to a coupling with the peak associated with a wagging of the bonds in the N–C imidazolium ring,39 and strong signals were also identified for all samples at ∼171 cm−1. The Raman spectra of the ILs are quite similar to others in the 287–476 cm−1 region, and at ∼1428 cm−1; both are associated with CH2(N)/CH3(N)CN relative to the imidazolium structural stretch, as has been previously reported.40 Low intensity peaks are identified in the 408 cm−1 region associated with CH2(N), CH3 and CH3(N)CH bonds.41 The O–C–O moiety has signals located at ∼588 cm−1 (relating to symmetric bending) and 1103 to 1328 cm−1 (relating to asymmetric bending).30 Less significant peaks, and common to all samples, are observed at 788–888 cm−1 arising from the ρ(CH2) rocking.42 Pronounced peaks at 1028–1054 cm−1, explained by the C–O symmetric strong peak,40 and overlapped with ν(C–C), are observed in the MC1–MC4 structures.
An interesting behaviour is identified in the 1554–1744 cm−1 region, related to CH3(N)HCH symmetric bending with CH2(N)/CH3(N)CN.40 For samples MC1 and MC4, these peaks disappear, while for samples MC2 and MC3, these peaks are closer and detached. This outcome can be explained by the similarities between the MC1/MC4 structures, which do not contain phenyl groups, and thus allow a greater flexibility in the organic chain. The relatively broad band ranging from 2800 to 2963 cm−1, which is assigned to symmetrical and asymmetrical C–H stretching (δ(CH3)) from the imidazolium cation (fingerprint bands), can serve as a useful probe to reflect structural change.43 The subtle difference may be associated with the elongation of the MC4 organic chain compared to the MC1 structure, whereas in samples with a phenyl structure (MC2/MC3), this peak is superimposed on the enlargement of nearby peaks. The band identified at ∼2950 cm−1 refers to νas(CH3) and was identified in all samples. This peak is due to the presence of several CH3 terminals in the structures.44 Another signal common to all samples occurs at 3078–3089 cm−1, attributed to ν(CH), and at ∼3096 cm−1, assigned to ν((C–H)ring.44 Signals identified at 3159 cm−1 (MC1), 3168 cm−1 (MC4), and 3178 cm−1 (MC2/MC3) are associated with the HC
CH ring, and ν(C–H).39Fig. 2 shows a strong band centred at 2959 cm−1, together with a small peak seen as a shoulder located at approximately 3433 cm−1. These peaks are associated with water molecules that interact preferentially with the anions, Br−⋯HOH⋯Br−.34,45 We compared the results obtained with ionic liquids used as ligands, ionogels and lanthanide salts (Fig. S15–S22, ESI†). On the whole, in the luminescent materials we observed, all signals referring to the structures of the ionic liquids besides the emission of some electronic transitions to the trivalent ions, are already identified in the Raman lines.46 The intensities of the characteristic peaks of the C
O bond, strong in the IL, became weaker in the ionogels, accompanied by a shift to a higher frequency. This outcome might have resulted from coordination of the IL with the metal in this moiety.47 Raman characterization of the Eu3+/Gd3+-ionogels is reported in detail in the ESI.† The assignments of several distinct Raman peaks of ionic liquids (MC1–4) and Gd3+/Eu3+ ionogels are listed in Tables S4 and S5 (ESI†), respectively.
The UV-Vis spectra of the ionogels present similar behaviour to that observed in the ILs (Fig. S26, ESI†). The bands identified at ∼200 nm and ∼280 nm (π; n → π* transitions) refer to the imidazolium moiety and intermolecular interactions.50 The complexes Eu(MC2)3(H2O)2 and Eu(MC3)3(H2O)2 show bands at ∼237 nm associated with absorption of the benzene ring. In comparison with the free ligand, these peaks were displaced by 10 nm (for the MC2 complex) and 17 nm (for the MC3 complex) to regions of lower energy. Fig. S26 (ESI†) shows the absorption spectra for ionogels with Eu3+. It's noteworthy that the ionogels are transparent in the visible range.
Steady-state photoluminescence spectra (Fig. 4) of the ionic liquids were obtained from pure samples and without the use of solvents. As observed in previous studies,50 the excitation spectra of the ILs confirm the observation of the prominence of intermolecular interactions, with a main excitation maximum at ∼323 nm and a very small component at ∼300 nm, the contribution of which grows with increasing concentration. The excitation spectra reached a maximum approximately at 358 nm, 332 nm, and 339 nm, for MC1, MC2 and MC4 compounds, respectively. This excitation is related to the π → π* transition. The emission for a single excitation wavelength is found to consist of two components, as previously reported.49 Ionic liquids, in particular those based on the imidazolium cation, should be more accurately considered as three dimensional networks of anions and cations linked together by weak interactions,36 as recorded using UV-Vis/NMR measurements. The short wavelength emission (here, at ∼430 nm) is due to the monomeric form of the imidazolium ion, and the long wavelength emission (∼550 nm) is due to its well-known associated forms.36 Emissions are identified as intraligand transitions and assigned as π* → π. The Stokes displacements were calculated to be: 79 nm for MC1, 104 nm for MC2, 96 nm for MC3, and 86 nm for MC4. We observed that the ionic liquids that have a benzene ring in their structure demonstrate a greater separation between the triplet levels, which causes a smaller overlap between the excitation–emission bands, and thus, a greater Stokes displacement.
The emission identified for the Gd3+ complexes corresponds to the emission of the ionic liquid because there is no transfer of energy from the IL to the Ln3+ ion. The Gd3+ emission (from the first excited state 6P7/2) has a higher energy, typically 32200 cm−1, corresponding to 310 nm.52 Emission and excitation spectra of the Gd3+–IL complexes are shown in Fig. S27–S30 (ESI†). In the comparison between these spectra and the photoluminescence results of the ionic liquids (MC1–4), it is noteworthy that there was a shift between the excitation and emission bands to lower wavelengths. The variation of the maximum emission (ΔλEm) and excitation (ΔλEx) was calculated, for each complex, indicating the modification of the ligand before coordination with the gadolinium. Ligand triplet state energy was estimated from the luminescence spectra of Gd(MCx)3(H2O)y complexes using the methods previously reported.29 Triplet level values and spectrum modifications (ΔλEm and ΔλEx) are set out in Table 1.
Complex | ΔλEx (nm) | ΔλEm (nm) | Triplet level energy of IL MC1–4 (cm−1) |
---|---|---|---|
Gd(MC1)3(H2O)3 | 24 | 16 | 23![]() |
Gd(MC2)3(H2O)2 | 11 | 38 | 24![]() |
Gd(MC3)3(H2O)2 | 21 | 54 | 26![]() |
Gd(MC4)3(H2O)3 | 16 | 39 | 25![]() |
We compare the diagrams of the singlet and triplet levels of the ILs and of Eu3+ in Fig. 5. All the systems showed that the triplet level of the ionic liquid is between 23000 and 30
000 cm−1. This fact is of immense relevance in the elaboration of high performance luminescent systems, due to the formation of high energy phonons, which deactivate the non-radiative ions.46 Thus, the ligands reported in this work present a remarkable potential in synergistic coordination with the Eu3+ ion. This synergy is discussed below.
Excitation spectra were obtained by monitoring the 5D0 → 7F2 line (619 nm for Eu(MC1)3(H2O)3; 614 nm for Eu(MC2)3(H2O)2; 616 nm for Eu(MC3)3(H2O)2; and 617 nm for Eu(MC4)3(H2O)3). The excitation spectrum consists of a series of sharp absorption lines, positioned in the 320–580 nm region, which can be attributed to transitions within the 4fn configurations of Eu3+ ions.10 The most intense transition of the excitation spectrum indicates the wavelength at which the direct excitation in the ion should give the greatest contribution to the photoluminescence of the compound. This wavelength is equal to 394 nm for Eu(MC1)3(H2O)3, Eu(MC2)3(H2O)2 and Eu(MC4)3(H2O)3, and 395 nm for Eu(MC3)3(H2O)2. Fig. 6, and Fig. S31–S34 (ESI†) show the corresponding emission spectra (red line) of the complexes obtained from direct ion excitation in Eu3+ (transition 7F0 → 5L6). In these spectra, five emission bands can be seen, centred at 580, 595, 612, 657 and 702 nm, distinguished in all complexes, and assigned to the 5D0 → 7FJ (J = 0–4) transitions in Eu3+.53
![]() | ||
Fig. 6 Excitation (black line; λem = 614 nm) and emission (red line; λex = 394 nm) of Eu(MC2)3(H2O)2. All measurements were performed in the steady state and at room temperature. |
The emission spectrum (Fig. S35, ESI†) obtained from the excitation in the band of each ligand (358 nm for Eu(MC1)3(H2O)3; 332 nm for Eu(MC2)3(H2O)2; 325 nm for Eu(MC3)3(H2O)2; and 340 nm for Eu(MC1)3(H2O)3) does not show the ligand phosphorescence band between 392–450 nm. Excitation of the complexes in the ligand part (MC1–4) gives the same spectral characteristics as the ion-centred Eu3+ emission. Emission spectra of Eu(MC1)3(H2O)3 show transition 0–4 with relevant intensity, which can be explained by the greater polarizability of the MC1 ligand compared to the other ILs (MC2, MC3 and MC4).54 The presence of donor groups or acceptors in the ligand structures modifies the electron density around the ion and interferes with the position of the 5D0 → 5F0 transition.49 This transition was observed at 579 nm for Eu(MC1)3(H2O)3, Eu(MC3)3(H2O)2 and Eu(MC4)3(H2O)3, and at 577 nm for Eu(MC2)3(H2O)2. This outcome suggests that MC2 presents a lower polarizability when compared to the other ligands. The lower intensity of the 0–4 transition of Eu(MC2)3(H2O)2 corroborates this fact. Less polarizable environments, in general, contribute to decreasing the covalent character of the metal–ligand bonds in the complexes (nephelauxetic effect), which implies a lower displacement for the 0–0 transition, and a lower intensity for the 0–4 transition.48 The presence of the 5D0 → 5F0 transition indicates that the Eu3+ ion is located in a symmetry site of the type Cs, Cn, or Cnv. This fact, together with the mono-exponential profile of each mono-exponential system of the 5D0 excited state radiative decay curve of the Eu3+ ion (Fig. S36 and S37, ESI†), indicates that Eu3+ ions predominantly exhibit coordination environments with only one symmetry centre.55 Emission spectra of the complexes (Fig. S31–S34, ESI†) show the multiplicities of the transitions, indicating that the point group for the environment of symmetry of the europium ion is around C2v.55 This outcome corroborates the hypothesis, represented by the 5D0 → 5F0 transition, that the chemical environment around Eu3+ must be of low symmetry. The fluorescence lifetimes were analysed throughout the spectral range of the europium ion (excitation at 7F0 → 5L6 and emission at 5D0 → 7F2), as shown in Fig. S36 (ESI†). We also examined the decay profiles measuring the temporal behaviour of the fluorescence by exciting the organic part of the samples and monitoring the fluorescence emission at 5D0 → 7F2 (Fig. S37, ESI†). All fluorescence decay profiles could be best fitted to a mono-exponential decay function, as shown in Fig. S36 and S37 (ESI†). The fluorescence lifetimes reported in this work are four times greater than those of luminescent gels, elaborated with europium and carboxylate ligands, recognized as high emission performance gels reported in the literature.56
We calculated the parameters of the experimental intensities: Ω2 and Ω4, Einstein spontaneous emission coefficient (Arad), the non-spontaneous emission coefficient (Anrad), quantum emission efficiency (η), and finally, R0–2/0–1 and R0–4/0–1, following well-established models in the literature for the determination of the energy transfer of ligand–lanthanide,29 the position and nature of excited states,57 and the intensity parameters.58Table 2 shows these results and shows the values of emission lifetimes from 5D0 in Eu3+.
The disparity between non-radiative rates can be explained by observing the vibrational spectra in the infrared region of these complexes. The FTIR spectra of Eu(MC1)3(H2O)3 and Eu(MC4)3(H2O)3 complexes show O–H (water-group) absorption bands at 3384 and 3393 cm−1, and the Eu(MC2)3(H2O)2 and Eu(MC3)3(H2O)2 complexes show lower intensity signals at 3476 and 3424 cm−1, respectively. The bands for the MC2/MC3 complexes are in a region of lower energy with respect to the bands for the MC1/MC4 complexes. The energy difference between 5D0 and 7F6 (for Eu3+ ion) is ∼12300 cm−1. Therefore, there is a better resonance condition involving three phonons (3 × 3393 cm−1) in the case of the Eu(MC1)3(H2O)3/Eu(MC4)3(H2O)3 complexes, causing a considerable increase in the rate of non-radiative decay. As can be seen in Table 2, MC1/MC4 complexes show lower efficiency. TGA results for Eu(MC1)3(H2O)3/Eu(MC4)3(H2O)3 showed 3 coordination water molecules, whereas for Eu(MC2)3(H2O)2 and Eu(MC3)3(H2O)2, only 2 water molecules were observed, which corroborates the vibrational spectra (FTIR/Raman) and justifies the higher quantum efficiency presented by these complexes. An analysis of Table 2 can be understood from the separation of the structures of the ligand. For ligands with alkyl chains (Eu(MC1)3(H2O)3 and Eu(MC4)3(H2O)3), we observe a high nonradiative decay rate in relation to the radiative decay rate, the lowest quantum efficiency, and the shortest lifetime. These observations can be associated with the high incidence of non-radiative processes, as vibrational coupling of O–H oscillators from water molecules coordinated with Eu3+. Eu(MC2)3(H2O)2 and Eu(MC3)3(H2O)2 had low nonradiative decay rates (Anrad) in relation to high radiative decay rates (Arad), higher quantum efficiency, and a lifetime beyond 1 ms. The Arad values and lifetimes are the highest reported compared to luminescent ionogels with europium ions reported in the literature.59–61 The values reported for the photophysical properties of Eu(MC2)3(H2O)2 and Eu(MC3)3(H2O)2 can be explained by the position of the triplet level in the energy diagram of the ionic liquids being resonant with the emitting level of Eu3+ (Fig. 5). We observed that Eu(MC2)3(H2O)2 showed greater efficiency compared to complexes with ligand structures similar to those proposed in this work.62 We compared the luminescence properties of different europium complexes to evaluate the effect of the alkyl chain length and flexibility. The ligand structures of MC1 and MC4 did not present steric impediments, besides having a greater structural flexibility, which implies a greater absorption of water and a closer approximation between neighbouring ions. The self-quenching process between neighbouring europium complexes could cause the nonradiative dissipation of excited states.13 The effective isolation and confinement of europium complexes inside more rigid structures, such as Eu(MC2)3(H2O)2 and Eu(MC3)3(H2O)2, were crucial to increasing luminescence efficiency, which explains the higher values of quantum efficiency and lifetime associated with these systems. Due to the nature of the chemical environment in terms of the intensity of 5D0 → 7F2 and 5D0 → 7F1 transitions, their intensity ratio (R0–2/0–1) could be used to assess the asymmetry of the Eu3+ coordination environment.63 A relatively higher ratio usually denotes a relatively higher degree of asymmetry and better luminescent monochromaticity.13 The disparity in the values found for R0–2/0–1 and R0–4/0–1 represents the structural differences caused by the ionic liquids used as ligands. The high values of R0–2/0–1 can be explained by the ionic nature of the ligands that involves a load compensation through the coordination network. The fact that no load transfer band in the spectrum can be seen supports this explanation. The high values of R0–2/0–1 suggest environments with low symmetry, as registered by the emission spectrum.
Complex | Ω 2 (10−20) | Ω 4 (10−20) | A rad (s−1) | A nrad (s−1) | η (%) | R 0–2/0–1 | R 0–4/0–1 | τ (ms) ± 0.001 ms |
---|---|---|---|---|---|---|---|---|
Eu(MC1)3(H2O)3 | 5.55 | 5.91 | 377.0 | 1377.9 | 21.5 | 3.16 | 1.52 | 0.5699 |
Eu(MC2)3(H2O)2 | 10.59 | 6.1 | 473.9 | 518.8 | 47.7 | 5.95 | 1.56 | 1.0072 |
Eu(MC3)3(H2O)2 | 8.61 | 7.78 | 451.4 | 524.4 | 46.3 | 4.84 | 1.98 | 1.0249 |
Eu(MC4)3(H2O)3 | 7.35 | 7.35 | 393.3 | 1082.0 | 26.7 | 4.15 | 1.87 | 0.6779 |
The Judd–Ofelt parameter Ω2 reported for Eu(MC2)3(H2O)2 is similar to the values found in the literature for organic ligands with benzenes and carboxylates.63,64 The disparity of the results reported in the literature compared to the other structures described in this work (Eu(MC1)3(H2O)3, Eu(MC3)3(H2O)2 and Eu(MC4)3(H2O)3) indicates the strong differences in the chemical environments for Eu3+ ions between these compounds. These differences can be explained by the benzene structure in the main chain (MC2) and benzenehexa carboxylic acid.63 In MC3, the benzene ring is only a branch of the main chain. In Eu(MC1)3(H2O)3 and Eu(MC4)3(H2O)3, which show great differences between Ω2 values, there is no benzene ring. The parameter Ω4 reflects the stiffness of the chemical environment around the lanthanide ion. The lowest Ω4 values were exhibited by the coordination network Eu(MC3)3(H2O)2, reflecting the network effect on the structural rigidity of the material. Similar ligand structures, Eu(MC1)3(H2O)3 and Eu(MC4)3(H2O)3, show congruent results that indicate the large size of the structural difference caused by the absence of the benzene ring. Preliminary measures of absolute emission quantum yields are reported in Table S8 (ESI†). We calculated the sensitization efficiency for the ligand based on previously reported methods.65 The high quantum yield observed for these complexes seems to be due to the calculated high yield energy transfer from these ligand states to the quasi-resonant Eu3+ energy levels.
The photostability of the Eu3+-complexes was evaluated by monitoring the emission/excitation spectra intermittently under continuous exposure to UV irradiation. Under UV exposure (UVA at 365 nm and UVB at 320 nm), the luminescence intensities of the intra configurational Eu3+ transitions suffer insignificant changes, as observed in Fig. S38 and S39 (ESI†). Using the intensity of the 5D0–7F2 transition, we produced a graph that provides evidence of the photostability of the systems against UV radiation (Fig. 7).
We studied the photostability by comparing the measurements performed on the same sample over three years (2016, 2017 and 2018). The emission and excitation spectra are reported in Fig. S41 (ESI†) for all the coordination systems. The Eu complexes showed great chemical stability and photostability to produce exactly the same spectral emissions/excitations even three years after synthesis of the ionogels.
The computational method24 employed in this work has already been well established in the literature and its main objective has been to provide data for a detailed study of the photoluminescence of for ionogels,25 crystals,26 composities27 and general complexes.28 This methodology presented consistent results for each of the elaborated complexes. All information pertaining to each structure and the geometry optimization method are reported in the ESI,† (Table S15). In the determination of the coordination sphere of the complexes, we found very reliable results for which we have the lowest percentage errors that have been reported for this computational procedure.26,62 The structures of the ionogels were optimized using the Sparkle/PM7 model. The numbers of water molecules are in accordance with those predicted by thermogravimetric analysis. Fig. 8 shows a representative coordination sphere for Eu(MC2)3(H2O)2. Figures of coordination geometries for Eu(MC1)3(H2O)3, Eu(MC3)3(H2O)2, and Eu(MC4)3(H2O)3 complexes are available in the ESI,† (Fig. S38–S41). As shown in Fig. 8, the ligand coordination comes from its carboxylate group in a chelating coordination, as predicted by FTIR analysis. The coordination polyhedron is organized in a distorted symmetry as a bicapped trigonal prism, which indicates that the arrangement is in point group symmetry C2v.66 This symmetry requires sp3d3f hybrids that are common coordination polyhedra for a coordination number of eight or nine.66 The proposed symmetry is consistent with the estimate from the emission spectra of all compounds. The distances between the metal centre and the oxygen atoms of the ionic liquid and coordinated water are shown in Tables S9–S11 in the ESI.†
![]() | ||
Fig. 8 Calculated geometry of the ground-state for Eu(MC2)3(H2O)2 using the PM7 model. (A) The coordination environment of the Eu3+ ion and (B) its coordination polyhedron. |
Spectroscopic properties as intensity parameters Ωλ (λ = 2, 4 and 6), energy transfer (WET) and back transfer (WBT) rates, radiative (Arad) and nonradiative (Anrad) decay rates, the quantum efficiency (η) and quantum yield (q) of europium compounds were calculated using semi-empirical models specialized for the study of photoluminescence of europium ions (LUMPAC).14 All results are exhibited in Table 3 and are in agreement with those obtained experimentally, evidencing the accuracy of the computational methodology. The outstanding photophysical properties inspired us to investigate the mechanisms of population transfer involved in the systems. Intramolecular energy transfer (ET) and back transfer (BT) were calculated, considering that the Eu3+ levels arose from the metal ion at an intermediate coupling.24 Theoretical data are reported for all complexes in Tables S12–S15 (ESI†). Energy transfer for Eu(MC1)3(H2O)3, Eu(MC2)3(H2O)2 and Eu(MC4)3(H2O)3 complexes follows a multipolar mechanism. For Eu(MC3)3(H2O)2, the profile suggests that the energy transfer occurs by an exchange mechanism.67,68 The population diagrams of the complexes showed the synergy between the carboxylate ionic liquids and the Eu3+ ions (Fig. S50–S54, ESI†).
Complex | Ω 2 (10−20 cm2) | Ω 4 (10−20 cm2) | A rad (s−1) | A nrad (s−1) | η (%) | R 0–2/0–1 | R 0–4/0–1 |
---|---|---|---|---|---|---|---|
Eu(MC1)3(H2O)3 | 6.59 | 4.02 | 308.4 | 1446.3 | 17.6 | 4.01 | 1.23 |
Eu(MC2)3(H2O)2 | 10.7 | 5.98 | 461.2 | 531.6 | 46.4 | 6.52 | 1.84 |
Eu(MC3)3(H2O)2 | 9.89 | 5.01 | 422.5 | 553.3 | 43.3 | 6.02 | 1.54 |
Eu(MC4)3(H2O)3 | 8.55 | 5.41 | 388.3 | 1086.8 | 26.3 | 5.20 | 1.67 |
We note that Eu3+/Gd3+ ionogels presented a similar coordination mode (as evidenced by FTIR and Raman spectra) and thermal stability until 223 °C (Fig. S23, ESI†) associated with high viscosity at room temperature. The interesting features of the soft materials include outstanding ionic conductivity due to a high content of ILs, easy coating on surfaces, and excellent luminescence properties (e.g., long lifetime, high colour purity, highlighted quantum efficiency, and photostability). These features might render them extremely valuable for various optical applications. The ionogels can also be entirely solution-processed, which is largely favourable to the development of low-cost and large-area lighting, or yet applicable for the composition of flexible displays in light emitting devices (LEDs), light-emitting electrochemical cells, and luminescent solar concentrators.
Our study demonstrates a complete analysis of the structure–property relationship for each complex (Eu(MC1)3(H2O)3, Eu(MC2)3(H2O)2, Eu(MC3)3(H2O)2 and Eu(MC4)3(H2O)3). The coordination polyhedron is in bicapped trigonal prism symmetry, which is consistent with emission spectrum estimates from all the compounds. The supramolecular ionogels had high viscosity at room temperature, transparency, and thermal stability until 223 °C, and strong photoluminescence (τ > 1 ms, η = 48% and φ = 27%). The extraordinary photophysical properties (presented in a detailed experimental and theoretical study) thus make these materials extremely valuable for optical applications such as light-emitting electrochemical cells. Our future research will focus on the fabrication of emitting devices based on Eu(MC1)3(H2O)3, Eu(MC2)3(H2O)2, Eu(MC3)3(H2O)2 and Eu(MC4)3(H2O)3 as thin films.
Footnote |
† Electronic supplementary information (ESI) available: Details of nuclear magnetic resonance, FTIR/Raman spectroscopy, elemental/thermophysical analysis, UV-Vis and photoluminescence spectrophotometry. See DOI: 10.1039/c8tc00658j |
This journal is © The Royal Society of Chemistry 2018 |