Next Article in Journal
Evaluation of Cr(VI) Removal from Tanning Effluents Using Magnetic Nanoparticles of Fe3O4 Synthesized with Olea europaea Bone Extract
Previous Article in Journal
Development of Fluorescent Chemosensors for Calcium and Lead Detection
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Microcrystalline Luminescent (Eu1-xLnx)2bdc3·nH2O (Ln = La, Gd, Lu) Antenna MOFs: Effect of Dopant Content on Structure, Particle Morphology, and Luminescent Properties

by
Stefaniia S. Kolesnik
1,
Nikita A. Bogachev
1,
Ilya E. Kolesnikov
1,
Sergey N. Orlov
1,2,
Mikhail N. Ryazantsev
1,3,
Gema González
4,
Mikhail Yu. Skripkin
1 and
Andrey S. Mereshchenko
1,*
1
Saint-Petersburg State University, 7/9 Universitetskaya Emb., 199034 St. Petersburg, Russia
2
Institute of Nuclear Industry, Peter the Great St. Petersburg Polytechnic University (SPbSU), 29 Polytechnicheskaya Street, 195251 St. Petersburg, Russia
3
Nanotechnology Research and Education Centre RAS, Saint Petersburg Academic University, ul. Khlopina 8/3, 194021 St. Petersburg, Russia
4
School of Physical Sciences and Nanotechnology, Yachay Tech University, Urcuqui 100119, Ecuador
*
Author to whom correspondence should be addressed.
Molecules 2024, 29(2), 532; https://doi.org/10.3390/molecules29020532
Submission received: 20 December 2023 / Revised: 10 January 2024 / Accepted: 17 January 2024 / Published: 21 January 2024
(This article belongs to the Special Issue Rare Earth Based Luminescent Materials)

Abstract

:
In this work, three series of micro-sized heterometallic europium-containing terephthalate MOFs, (Eu1-xLnx)2bdc3·nH2O (Ln = La, Gd, Lu), are synthesized via an ultrasound-assisted method in an aqueous medium. La3+ and Gd3+-doped terephthalates are isostructural to Eu2bdc3·4H2O. Lu3+-doped compounds are isostructural to Eu2bdc3·4H2O with Lu contents lower than 95 at.%. The compounds that are isostructural to Lu2bdc3·2.5H2O are formed at higher Lu3+ concentrations for the (Eu1-xLux)2bdc3·nH2O series. All materials consist of micrometer-sized particles. The particle shape is determined by the crystalline phase. All the synthesized samples demonstrate an “antenna” effect: a bright-red emission corresponding to the 5D0-7FJ transitions of Eu3+ ions is observed upon 310 nm excitation into the singlet electronic excited state of terephthalate ions. The fine structure of the emission spectra is determined by the crystalline phase due to the different local symmetries of the Eu3+ ions in the different kinds of crystalline structures. The photoluminescence quantum yield and 5D0 excited state lifetime of Eu3+ are equal to 11 ± 2% and 0.44 ± 0.01 ms, respectively, for the Ln2bdc3·4H2O structures. For the (Eu1-xLux)2bdc3·2.5H2O compounds, significant increases in the photoluminescence quantum yield and 5D0 excited state lifetime of Eu3+ are observed, reaching 23% and 1.62 ms, respectively.

Graphical Abstract

1. Introduction

Light-emitting materials based on the metal–organic frameworks (MOFs) of rare earth element (REE) ions have been widely discussed in recent research works because of the wide area of applications, ranging from LEDs and different sensors to bioimaging materials and medicine [1,2,3,4]. For practical application, the particle size of such compounds is one of the most important factors. Small-sized particles (nano- and micro-sized) have a large specific surface area, which provides high sorption properties to compounds and allows them to be used as sensor materials for heavy metal ion detection [3,5,6,7,8]. Such highly dispersed MOFs can be obtained via several synthetic approaches, namely coprecipitation [9], hydrothermal [10], solvothermal [11], sol–gel [12], magnetic-field-induced [13], and some other synthetic methods. One of the simplest, most efficient, and low-cost approaches for the synthesis of small-sized MOF particles is the ultrasound-assisted method, which was recently used to obtain small-sized, REE-based, MOF Eu2bdc3·4H2O nanoparticles, with average particle sizes of 8 ± 2 nm [14]. These nano-sized particles were shown to be one of the most sensitive luminescent MOF-based sensors for Cu2+, Cr3+, and Fe3+ ions. This method of synthesis allows for the preparation of microparticles with certain morphologies and sizes. The ultrasound-assisted method is quite unpopular today but has big prospects because of its advantages in simplicity of use and the ability to control the physical parameters of the target product.
Since the probability of f-f transitions is very small (Laporte’s rule), direct excitation into the 4f excited levels of ions is unfavorable, which limits the application of such luminescent materials. One of the useful approaches to overcome this problem is the so-called “antenna effect”, which suggests using light-harvesting ligands that transfer absorbed energy onto lanthanide ions [15,16]. Typical ligands used in such antenna complexes are aromatic and unsaturated molecules, like calixarenes, bipyridines, phenanthroline derivatives, and carboxylates, including terephthalates [12,17,18,19,20].
The mutual presence of both luminescent and non-luminescent rare earth element (REE) ions can significantly affect the photophysical properties of these compounds. In the mixed terephthalates and diphenylmethionates Eu-Y, Eu-Gd, Tb-Y, and Tb-Gd, it has been shown that the quantum yield and luminescence lifetime significantly depend on the concentration of the non-luminescent ions yttrium and gadolinium [12,17,18], but the reasons that determine the observed effects have practically not been discussed. The number of works reporting the effect of MOF co-doping using non-luminescent REEs is limited, and these works are non-systematic. For example, we have not found any works that have considered the effect of co-doping with lutetium and lanthanum ions (they do not exhibit luminescence) in contrast to gadolinium, which is commonly used as a dopant. In addition, in the works devoted to the study of heterometallic REE-based MOFs containing both luminescent and non-luminescent ions, the mechanism of the mutual effect of REE on the photophysical properties is practically not disclosed. Recently, we reported on heterometallic europium–lutetium and terbium–lutetium terephthalate metal–organic frameworks (MOFs) that demonstrate a strong correlation between the luminescent properties of the complexes and the crystalline structure [19,20]. It has been shown that the luminescence of lanthanide (III) ions strictly depends on the local symmetry of emitting lanthanide ions [19,20,21,22,23,24]. Therefore, in our studies, we have found that the fine structure of luminescence spectra, the lifetimes of excited states, and quantum yields change with crystalline phase change upon varying REE ion contents. For example, in a series of (TbxLu1-x)2bdc3·nH2O compounds (bdc—1,4-benzenedicarboxylate) containing more than 30 at.% of Tb3+, only one crystalline phase is formed, Ln2bdc3·4H2O. At lower Tb3+ concentrations, terephthalates crystallize as a mixture of Ln2bdc3·4H2O and Ln2bdc3·10H2O or Ln2bdc3; the exact composition of the product depends on the reagent concentrations used in the synthesis. The 5D4 excited state Tb3+ lifetimes and photoluminescence quantum yield (PLQY) of the anhydrous phase are significantly larger than the corresponding values for terephthalate tetrahydrates and decahydrates. In the case of europium-based compounds, we have found similar results: the luminescence quantum yield of Eu3+ was significantly larger for Eu-Lu terephthalates containing a low concentration of Eu3+ due to the absence of Eu-Eu energy transfer and the presence of the anhydrous Ln2bdc3 crystalline phase, with a significantly smaller nonradiative decay rate compared to the Ln2bdc3·4H2O. Therefore, we concluded that the luminescence of Eu3+ and Tb3+ is quenched by the coordinated water molecules.
To continue our study of the effect of non-luminescent heavy atoms on the interrelated structural and luminescent properties of heterometallic lanthanide-based terephthalates, in the current work, we present the results of our study on microcrystalline europium-based terephthalates doped with lutetium(III), lanthanum(III) and gadolinium(III). In our work, we obtained such MOFs using ultrasound-assisted methods. This unique, simple method allowed us not only to obtain small particles but also to achieve a reproducible synthesis of particles with the desired properties and certain morphology and size. Furthermore, using the ultrasound-assisted method, we obtained a previously unknown crystalline compound, namely Lu2bdc3·2.5H2O, and the isostructural heterometallic Eu-Lu terephthalates, (Eu1-xLnx)2bdc3·2.5H2O.

2. Results and Discussion

2.1. Crystalline Structure

In Figure 1a,b, the PXRD patterns of the synthesized MOFs (Eu1-xLnx)2bdc3·nH2O (Ln = La, Gd) are shown. We found that mixed Eu-La and Eu-Gd terephthalates were isostructural to the Ln2bdc3·4H2O crystalline phase (Ln = Ce − Yb) [25] at the whole concentration range, and additional peaks were not observed. In the Ln2bdc3·4H2O structure, the lanthanide (III) ions are bound to two water molecules and six terephthalate ions through oxygen atoms, and the Ln3+ coordination number (CN) is equal to 8, Figure 1c. The PXRD patterns of (Eu1-xLux)2bdc3·nH2O MOFs are shown in Figure 2. Analysis of the data demonstrates that the structure of the compounds formed depends on the lutetium content. Thus, compounds with a concentration of lutetium (III) ions 90 at.% and less are isostructural to the Ln2bdc3·4H2O crystalline phase. In contrast, when the concentration of lutetium is 95 at.% and more, the positions of the PXRD maxima do not correspond to any known Ln2bdc3·nH2O structure, such as Eu2bdc3·4H2O [25], Tb2bdc3 [25,26], Yb2bdc3·6H2O [27], Lu2bdc3·10H2O [28], Sc2bdc3 [29], Er2bdc3·3H2O [30].
To prove that compounds containing more than 95 at.% of lutetium have (Eu1-xLux)2bdc3·nH2O composition (containing some water molecules),samples were calcined at the temperature of 200 °C. The positions of the PXRD maxima of the obtained substances correspond to the anhydrous terephthalate structures (Eu1-xLux)2bdc3, which are isostructural to Tb2bdc3 (Figure 3). The formation of anhydrous terephthalate is the result of the loss of coordinated water molecules by (Eu1-xLux)2bdc3·nH2O, resulting in the formation of (Eu1-xLux)2bdc3. Thermogravimetric analysis of Lu2bdc3·nH2O in a temperature range of 30–230 °C allowed us to find the number of coordinated water molecules, n, in the synthesized microcrystalline lutetium terephthalate hydrate Lu2bdc3·nH2O (Figure 4a) under the assumption that this compound was formed in a single crystalline phase. The mass loss was observed at 60–150 °C. As previously reported [19,20], the mass loss in this temperature range can be attributed to the dehydration of the compounds, resulting in the formation of anhydrous terephthalate: Lu2bdc3·nH2O → Lu2bdc3 + nH2O. The average weight loss was found to be (5.31 ± 0.18)%, which corresponds to 2.62 ± 0.09 water molecules per unit formula of Lu2bdc3·nH2O. The product of calcination was assigned to the anhydrous terephthalate based on the PXRD data showing that the calcination product, Lu2bdc3, is isostructural to Tb2bdc3, Figure 3. Therefore, we proposed that the new unknown crystalline phase of obtained lutetium terephthalate hydrate has the composition of Lu2bdc3·2.5H2O. This crystalline phase dominates at the lutetium content of 95 at.% and more in our samples. Thermogravimetric analysis of Eu2bdc3·4H2O (Figure 4b) confirmed the number of water molecules per unit formula of europium(III) terephthalate hydrate, where weight loss of 8.51% (5.31% ± 0.18) corresponds to 4.32 water molecules. Most probably, the thermogravimetric analysis gives overestimated values of the number of water molecules per unit formula of metal terephthalate because of the presence of a small amount of the absorbed water in the MOF pores.
Unit cell parameters were refined using UnitCell software [31], Table 1. This program can retrieve unit cell parameters from diffraction data using a least-squares method from the positions of the indexed diffraction maxima of PXRD patterns (Pawley method [32]). For the compounds (Eu1-xLax)2bdc3·4H2O, the increase in La3+ content leads to unit cell volumes increase due to a higher ionic radius of La3+ ions (1.160 Å, the coordination number is eight) than the ionic radius of Eu3+ ions (1.066 Å) [33]. The ionic radius of the Gd3+ ion (1.053 Å) is close to that of Eu3+. Therefore, the unit cell parameters do not change significantly in the (Eu1-xGdx)2bdc3·4H2O series. The Lu3+ ion (ionic radius is 0.977 Å) is smaller than Eu3+; therefore, substitution of Eu3+ by Lu3+ ion results in a decrease in the unit cell volumes in (Eu1-xLux)2bdc3·4H2O series with the lutetium (III) ions content of 90 at.% and less, where the MOF is formed as tetrahydrate, Ln2bdc3·4H2O.

2.2. Particle Morphology

Scanning electron microscopy (SEM) was used to observe the shape and the size of the particles of the synthesized MOFs (Eu1-xLnx)2bdc3·nH2O (Ln = La, Gd, Lu). The SEM images are shown in Figure 5, where (Eu1-xLax)2bdc3·nH2O, (Eu1-xGdx)2bdc3·nH2O, (Eu1-xLux)2bdc3·nH2O compounds are presented in the left, central, and right columns, respectively. The length and width of the particles obtained from the SEM images are given in Table 2. In the (Eu1-xLax)2bdc3·4H2O and (Eu1-xGdx)2bdc3·4H2O series, the particles have a similar oval plate shape, which can resemble one a petal or a leaf, as well as a similar size of about 6 × 2 μm. We noticed that the particles have different shapes depending on the content of lutetium ion in the (Eu1-xLux)2bdc3·nH2O series. Thus, at the Lu3+ content below 90 at.%, the particles have the shape of rods and a size of about 4 × 0.8 μm. However, in (Eu1-xLux)2bdc3·nH2O MOFs with a concentration of Lu3+ more than 90%, the particles are “brick”-shaped and are significantly larger, about 10 × 5 μm. At the Lu3+ concentration of 90 and 95 at.%, the mixture of “bricks” and “rods” is observed. The shape difference probably results from the change in the crystalline structure at the Lu3+ concentration of 95% among the (Eu1-xLux)2bdc3·nH2O series. Detailed analysis of the shape (Eu1-xLax)2bdc3·4H2O and (Eu1-xGdx)2bdc3·4H2O microparticles revealed that leaf-shaped microparticles are made of rods, whereas the separate rods about of the same size are observed in the (Eu1-xLux)2bdc3·4H2O (x = 0.8, 0.9) samples. We propose that the reason might be in different dopant properties. Lu3+ is the smallest ion among the trivalent lanthanide cations [33]. Therefore, the Lu3+ ion possesses the highest positive charge density among lanthanides. Therefore, we propose that the lutetium terephthalate surface has the highest surface charge density among the series. In the literature [34], a higher surface charge density corresponds to a higher colloidal stability of the solution. We believe that small rod”-shaped particles initially form in the mixed terephthalates containing 80–90 at.% dopant. Therefore, Gd and La-doped particles coagulate, forming “leaf”-shaped particles, whereas Lu-doped particles are still colloidally stable and save their “rod” form if they correspond to the Ln2bdc3·4H2O crystalline phase.
It is worth noting that the ultrasound-assisted method allowed us to obtain microparticles with a very small size variance (near 1–3 μm). This result suggests that using this method, it is possible to obtain particles with a certain morphology and size by changing the dopant content.

2.3. Luminescence Properties

Europium and terbium terephthalates demonstrate a pronounced antenna effect where terephthalate ion as a synthesizer or an “antenna” effectively absorbs UV radiation and transfers energy to a luminescent lanthanide ion followed by the metal-centered emission [18]. Upon the excitation, the terephthalate ion is promoted into the Sn state, followed by the fast internal conversion to the S1 state. Due to the heavy atom effect caused by the lanthanide atom, the S1 state efficiently undergoes intersystem crossing to the T1 triplet electronic excited state [18]. The T1 state of the terephthalate ion is close in energy to the 5D1 energy level of the Eu3+ ion. Therefore, an efficient energy transfer from sensitizer to luminescent lanthanide ion occurs. The 5D1 level of the Eu3+ then undergoes an internal conversion into the 5D0 state, followed by the emission into the 7FJ (J = 0–4) lower-lying energy levels.
For the synthesized luminescent materials (Eu1-xLnx)2bdc3·nH2O (Ln = La, Gd, Lu and x = 0, 0.80, 0.90, 0.95, 0.98), emission spectra were measured upon 310-nm excitation (Figure 6 and Figure 7). The emission spectra for the Gd and La-doped compounds are similar to the pure europium terephthalateemission spectrum. This observation agrees with the PXRD data showing that (Eu1-xLnx)2bdc3·nH2O (Ln = La, Gd) compounds are isostructural to Eu2bdc3·4H2O compound (Figure 6). The emission spectra of Gd and La-doped terephthalates consist of the narrow bands corresponding to 5D0-7FJ (J = 0–4) transitions of Eu3+: 5D0-7F1 (588.2 and 591.6 nm), 5D0-7F2 (614.6 nm), and 5D0-7F4 (696.6 nm). 5D0-7F0 and 5D0-7F3 bands were not observed in the emission spectra due to their weak intensity.
The emission spectra of the (Eu1-xLux)2bdc3·nH2O compounds are shown in Figure 7. This figure clearly shows the difference in the fine structure of the measured emission spectra for the Lu concentration of 0, 80, 90, 95, and 98 at.%. The first three emission spectra are identical to the spectrum of Eu2bdc3·4H2O. This observation is consistent with PXRD data demonstrating that (Eu1-xLux)2bdc3·nH2O compounds (x = 0.80, 0.90) crystalize in Eu2bdc3·4H2O phase without formation of any additional phase. The emission spectra of the (Eu1-xLux)2bdc3·nH2O compounds consist of narrow bands corresponding to 5D0-7F1 (587.9, 591.3, and 591.8 nm), 5D0-7F2 (614.2 nm), and 5D0-7F4 (696.7 nm). On the contrary, the emission spectra of the (Eu1-xLux)2bdc3·nH2O (x = 0.95, 0.98) compounds are different from the spectrum of pure europium terephthalate. Such difference can be explained by the formation of another phase, namely Ln2bdc3·2.5H2O, where the local symmetry of the europium (III) ion is different from that of Eu2bdc3·4H2O. These spectra consist of narrow bands corresponding to 5D0-7F1 (588.6, 591.6, and 593 nm), 5D0-7F2 (613.7, 616.0, and 622.0 nm), and 5D0-7F4 (698.2 nm).
The photoluminescence decay curves were measured under UV-excitation and monitored at 614 nm (5D0-7F2) for the (Eu1-xLnx)2bdc3·nH2O (Ln = La, Gd, Lu and x = 0, 0.80, 0.90, 0.95, 0.98) compounds (Figure 8 and Figure 9). The photoluminescence decay curves of La and Gd doped terephthalates (Figure 8) were accurately fitted by single exponential functions (Equation (1)), resulting in time constants τ of about 0.43–0.47 ms for the Gd doped terephthalates and about 0.43–0.45 ms for the La-doped terephthalates (Table 3). For the (Eu1-xLux)2bdc3·nH2O materials, the photoluminescence decay curves were fitted by the single exponential functions (Equation (1)) at the lutetium content less than 90 at.% resulting in time constants τ of about 0.43–0.46 ms (Figure 9). Meanwhile, at lutetium concentrations more than 90 at.% The photoluminescence decay curves can be fitted only using the double exponential functions (Equation (2)) with the time constants τ1 of about 0.34–0.46 ms and τ2 of about 1.5 ms. The two-exponential decay of the highly doped Eu-Lu terephthalates (Eu1-xLux)2bdc3·nH2O (x = 0.95, 0.98) indicates the two different coordination sites of the Eu3+ ion. The fast component (0.3–0.5 ms) is close to that of the water-coordinated Eu3+ ion in the Ln2bdc3·4H2O structure, whereas the slow component (1.5–1.6 ms) corresponds to the Eu3+ ion exclusively coordinated to carboxylic groups of terephthalate ions [20] due to the absence of luminescence quenching of Eu3+ by coordinated water molecules containing high-frequency OH vibrational modes.
For all synthesized materials, photoluminescence quantum yields (PLQYs) were measured (Table 3). PLQYs for the Gd and La-doped terephthalates are found to be approximately equal. But for the (Eu1-xLux)2bdc3·nH2O materials, PLQYs grow with the Lu concentration increase and reach a maximum of 23% at the 95 at.% Lu. The significantly higher values of photoluminescence quantum yield in Eu-Lu terephthalates at high Lu content can be explained by the presence in these materials’ Eu3+ coordination sites bound only to carboxylic groups of terephthalate ions [20] and not to the water molecules, which quenches the Eu3+ luminescence.
I = I 1 · e t τ
I = I 1 · e t τ 1 + I 2 · e t τ 2
Due to the different nature of the f-f transitions of Eu3+ ions, these ions can be used as structural probes [35,36]. Thus, the probability of hypersensitive 5D0-7F2 forced electric dipole transition is strongly affected by the local environment of the europium ion, whereas the probability of 5D0-7F1 magnetic dipole transition intensity is significantly less sensitive to changes in the Eu3+ coordination sphere. The analysis of asymmetry ratio Ras, which is equal to the ratio of the integral intensity of these transitions (5D0-7F2)/(5D0-7F1), allows one to track the changes in the local environment of the europium (III) ions [37,38]. The higher values of the asymmetry ratio correspond to the greater deviation of the Eu3+ environment from the centrosymmetric one [39]. The asymmetry ratio for the synthesized compounds is given in Table 4. The asymmetry ratio decreases with the increasing dopant concentration caused by the distortion of the crystal structure near the luminescent centers. The lutetium-doped materials have a minimal asymmetry ratio at the whole concentration range of the dopant, especially when the concentration of lutetium is 95 at.% and even more when the phase transition occurs. The reason for this effect is that the lutetium (III) ion has a smaller ionic radius than the gadolinium(III) and lanthanum(III) doping ions, and the environment of the luminescent center is not distorted. Lanthanum and gadolinium-doped terephthalates have nearly equal asymmetry ratios because of the almost identical structures of these materials.

3. Materials and Methods

3.1. Materials

Lutetium (III) chloride hexahydrate (>99%), gadolinium chloride hexahydrate (>99%), lanthanum chloride hexahydrate (>99%), and europium (III) chloride hexahydrate (>99%) were purchased from Chemcraft (Kaliningrad, Russia). Benzene-1,4-dicarboxylic (terephthalic, H2bdc) acid (>98%), sodium hydroxide (>99%), nickel (II) chloride hexahydrate (>99%), and EDTA disodium salt (0.1 M aqueous solution) were purchased from Sigma-Aldrich Pty Ltd. (Taufkirchen, Germany) and used without additional purification. The 0.3 M aqueous solution of disodium terephthalate (Na2bdc) was prepared by dissolving terephthalic acid in an aqueous sodium hydroxide solution and then diluted to obtain the 10 mM Na2bdc aqueous solution. The 0.2 M EuCl3 LuCl3, LaCl3, and GdCl3 solutions were prepared, standardized using back complexometric titration as described earlier [40], and then diluted to obtain the solutions containing the mixture of REE chlorides used in the synthesis of MOFs (Table 5).

3.2. Synthesis

In the current work, three series of microcrystalline heterometallic MOFs were synthesized, namely (Eu1-xLax)2bdc3·nH2O, (Eu1-xGdx)2bdc3·nH2O, and (Eu1-xLux)2bdc3·nH2O (x = 0, 0.80, 0.90, 0.95, 0.98, 1). The MOFs were synthesized by a slow dropwise addition of 10 mL of 10 mM Na2bdc aqueous solution to the 10 mL solution containing chlorides of the abovementioned lanthanides with a total REE concentration of 5 mM accompanied by ultrasonication (40 kHz, 60 W) and vigorous stirring. The initial concentrations of REE ions are given in Table 5. After the dropwise addition that took 5 min, the reaction mixture was kept in an ultrasonic bath for 10 more minutes. As a result, white precipitates of europium(III)-lanthanide(III) terephthalates were formed. The MOFs were separated from the reaction mixture by centrifugation (4000 g) and washed with distilled water five times. The resulting materials were dried under reduced pressure (0.02 atm.) at 20 °C in the desiccator for 1 day. As a result, heterometallic MOFs (Eu1-xLnx)2bdc3·nH2O (Ln = Lu, La, Gd and x = 0, 0.80, 0.90, 0.95, 0.98, 1) were obtained. Ratios of the rare earth elements in the synthesized compounds were confirmed by the EDX method (Table 6). EDX shows that, in general, the ratio of elements taken for synthesis is maintained during crystallization.

3.3. Methods

The relative content of the rare earth elements in the synthesized compounds was confirmed by energy-dispersive X-ray spectroscopy. The morphology of the particles was characterized by scanning electron microscopy (SEM) on a Zeiss Merlin electron microscope (Zeiss, Oberkochen, Germany) using an energy-dispersive X-ray spectroscopy (EDX) module (Oxford Instruments INCAx-act, High Wycombe, UK). Powder X-ray diffraction (PXRD) measurements were performed on a D2 Phaser (Bruker, Billerica, MA, USA) X-ray diffractometer using Cu Ka radiation (λ = 1.54056 Å). The thermal behavior of the compounds was studied by means of thermogravimetry using a Thermo-microbalance TG 209 F1 Libra (Netzsch, Selb, Germany) with a heat-up rate of 10 °C/min. To carry out photoluminescence studies, the synthesized samples (20 mg) and potassium bromide (300 mg) were pressed into pellets (diameter 13 mm). The luminescence spectra were recorded on Fluorolog-3 fluorescence spectrometer (Horiba Jobin Yvon, Kyoto, Japan). Lifetime measurements were performed using the same spectrometer using a pulsed Xe lamp (pulse duration 3 µs).

4. Conclusions

In this article, we reported on the morphology and the photoluminescence properties of the three series of microcrystalline heterometallic europium-containing terephthalate metal–organic frameworks synthesized by ultrasound-assisted method from diluted aqueous solutions: (Eu1-xLax)2bdc3·nH2O, (Eu1-xGdx)2bdc3·nH2O and (Eu1-xLux)2bdc3·nH2O (x = 0–1). The effect of the dopant concentration on the structural properties was revealed. Thus, the La3+ and Gd3+-doped terephthalates are isostructural to Eu2bdc3·4H2O, but the Lu3+ doped compounds are isostructural to Eu2bdc3·4H2O only when the Lu content is lower than 95 at.%; at higher Lu3+ content, the new structure namely Lu2bdc3·2.5H2O was obtained. The unit cell parameters were optimized, and the unit cell parameters were refined for the compounds corresponding to Ln2bdc3·4H2O crystalline phase. Unit cell parameters strongly depend on the content and ionic radius of the doping ion in the (Eu1-xLax)2bdc3·nH2O and (Eu1-xLux)2bdc3·nH2O series. Thus, the substitution of Eu3+ for larger La3+ ions increases the unit cell volumes, whereas doping by the smaller Lu3+ ion results in a decrease in the unit cell volume. This effect is more pronounced for a higher concentration of doping ions. The ionic radius of the Gd3+ ion is close to that of Eu3+. Therefore, the unit cell parameters do not change significantly in the (Eu1-xGdx)2bdc3·4H2O series. The analysis of the morphology of the synthesized materials by scanning electron microscopy has demonstrated that the ultrasound-assisted method results in the formation of particles that have a size of several micrometers and a shape determined by the crystalline phase. In the (Eu1-xLax)2bdc3·4H2O and (Eu1-xGdx)2bdc3·4H2O series, the particles have a similar “leaf” shape and size of approximately 6 × 2 μm. In the (Eu1-xLux)2bdc3·nH2O series, particles have different shapes depending on the content of lutetium ion. At the Lu3+ content below 90 at.%, which corresponds to the Ln2bdc3·4H2O crystalline phase, the particles have the shape of rods and a size of approximately 4 × 0.8 μm. However, in (Eu1-xLux)2bdc3·nH2O MOFs with a concentration of Lu3+ more than 90% (Ln2bdc3·2.5H2O crystalline phase), the particles are “brick”-shaped and significantly larger, about 10 × 5 μm. At the Lu3+ concentration of 90 and 95 at.%, the mixture of “bricks” and “rods” is observed.
All synthesized samples containing Eu3+ demonstrated a bright-red emission corresponding to the 5D0-7FJ (J = 1, 2, 4) transitions of Eu3+ ions upon 310-nm excitation to the singlet electronic excited state of terephthalate ions due to the “antenna” effect. The fine structure of the emission spectra is determined by the crystalline phase due to the different local symmetry of the Eu3+ ions in different types of crystalline structure, namely Ln2bdc3·4H2O and Ln2bdc3·2.5H2O. In the (Eu1-xLax)2bdc3·nH2O and (Eu1-xGdx)2bdc3·nH2O series, the photoluminescence quantum yields and 5D0 excited state lifetimes are equal to 11 ± 2% and 0.44 ± 0.01 ms, respectively, and almost do not depend on the content of La3+ and Gd3+. The substitution of Eu3+ for Lu3+ results in an increase in both the photoluminescence quantum yield (up to 23% at the Lu3+ content of 95%) and 5D0 excited state lifetime (up to 1.62 ms) in agreement with the phase transition from Ln2bdc3·4H2O to Ln2bdc3·2.5H2O. Therefore, in the current work, we have shown that the ultrasound-assisted method allows us to control the morphology of the particles and the luminescent properties of selected heterometallic antenna MOFs (Eu1-xLnx)2bdc3·nH2O (Ln = La, Gd, Lu) by changing the concentration of doping REE ion. This outstanding result will allow one to obtain particles with predetermined physical properties, morphology, and size. Furthermore, using the ultrasound-assisted method, we discovered the new previously unknown crystalline compound, namely Lu2bdc3·2.5H2O, and the isostructural heterometallic Eu-Lu terephthalates, (Eu1-xLnx)2bdc3·2.5H2O, which can be synthesized exclusively by our synthetic procedure under ultrasonic conditions.

Author Contributions

Conceptualization, A.S.M., N.A.B. and S.S.K.; methodology, A.S.M. and S.S.K.; validation, M.N.R. and I.E.K.; formal analysis, A.S.M., N.A.B. and S.S.K.; investigation, A.S.M., S.S.K., S.N.O., I.E.K. and M.Y.S.; resources, A.S.M., M.Y.S. and N.A.B.; data curation, A.S.M. and S.S.K.; writing—original draft preparation, A.S.M., N.A.B. and S.S.K.; writing—review and editing, M.Y.S., I.E.K., N.A.B., S.S.K., G.G. and A.S.M.; visualization, A.S.M. and S.S.K.; supervision, A.S.M.; project administration, A.S.M.; funding acquisition, A.S.M. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the Russian Science Foundation under grant no. 22-73-10040 (https://rscf.ru/en/project/22-73-10040/, accessed on 27 November 2023).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data presented in this study are available in the article.

Acknowledgments

The measurements were performed in the Research Park of Saint-Petersburg State University (Magnetic Resonance Research Centre, Chemical Analysis and Materials Research Centre, Cryogenic Department, Interdisciplinary Resource Centre for Nanotechnology, Centre for X-ray Diffraction Studies, Centre for Optical and Laser Materials Research, Thermogravimetric and Calorimetric Research Centre, and Centre for Innovative Technologies of Composite Nanomaterials).

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Younis, S.A.; Bhardwaj, N.; Bhardwaj, S.K.; Kim, K.-H.; Deep, A. Rare Earth Metal–Organic Frameworks (RE-MOFs): Synthesis, Properties, and Biomedical Applications. Coord. Chem. Rev. 2021, 429, 213620. [Google Scholar] [CrossRef]
  2. Neufeld, M.J.; Winter, H.; Landry, M.R.; Goforth, A.M.; Khan, S.; Pratx, G.; Sun, C. Lanthanide Metal–Organic Frameworks for Multispectral Radioluminescent Imaging. ACS Appl. Mater. Interfaces 2020, 12, 26943–26954. [Google Scholar] [CrossRef] [PubMed]
  3. Xia, N.; Chang, Y.; Zhou, Q.; Ding, S.; Gao, F. An Overview of the Design of Metal-Organic Frameworks-Based Fluorescent Chemosensors and Biosensors. Biosensors 2022, 12, 928. [Google Scholar] [CrossRef] [PubMed]
  4. Yu, Z.; Kang, S.; Tai, M.; Wang, J.; Tian, Q.; Jin, D.; Wang, L. Synthesis, Modulation, and Characterization of Ln3+ Ions Doped Metal−organic Frameworks for WLED Applications. Dyes Pigments 2023, 209, 110897. [Google Scholar] [CrossRef]
  5. Zhao, S.-N.; Wang, G.; Poelman, D.; Voort, P. Luminescent Lanthanide MOFs: A Unique Platform for Chemical Sensing. Materials 2018, 11, 572. [Google Scholar] [CrossRef] [PubMed]
  6. Dou, X.; Sun, K.; Chen, H.; Jiang, Y.; Wu, L.; Mei, J.; Ding, Z.; Xie, J. Nanoscale Metal-Organic Frameworks as Fluorescence Sensors for Food Safety. Antibiotics 2021, 10, 358. [Google Scholar] [CrossRef]
  7. Puglisi, R.; Pellegrino, A.L.; Fiorenza, R.; Scirè, S.; Malandrino, G. A Facile One-Pot Approach to the Synthesis of Gd-Eu Based Metal-Organic Frameworks and Applications to Sensing of Fe3+ and Cr2O72− Ions. Sensors 2021, 21, 1679. [Google Scholar] [CrossRef]
  8. Duan, L.; Zhang, C.; Cen, P.; Jin, X.; Liang, C.; Yang, J.; Liu, X. Stable Ln-MOFs as Multi-Responsive Photoluminescence Sensors for the Sensitive Sensing of Fe3+, Cr2O72− and Nitrofuran. CrystEngComm 2020, 22, 1695–1704. [Google Scholar] [CrossRef]
  9. Kolesnik, S.S.; Nosov, V.G.; Kolesnikov, I.E.; Khairullina, E.M.; Tumkin, I.I.; Vidyakina, A.A.; Sysoeva, A.A.; Ryazantsev, M.N.; Panov, M.S.; Khripun, V.D.; et al. Ultrasound-Assisted Synthesis of Luminescent Micro- and Nanocrystalline Eu-Based MOFs as Luminescent Probes for Heavy Metal Ions. Nanomaterials 2021, 11, 2448. [Google Scholar] [CrossRef]
  10. Yin, H.-Q.; Wang, X.-Y.; Yin, X.-B. Rotation Restricted Emission and Antenna Effect in Single Metal–Organic Frameworks. J. Am. Chem. Soc. 2019, 141, 15166–15173. [Google Scholar] [CrossRef]
  11. Wu, J.-Q.; Ma, X.-Y.; Liang, C.-L.; Lu, J.-M.; Shi, Q.; Shao, L.-X. Design of an Antenna Effect Eu(III)-Based Metal–Organic Framework for Highly Selective Sensing of Fe3+. Dalton Trans. 2022, 51, 2890–2897. [Google Scholar] [CrossRef] [PubMed]
  12. Alammar, T.; Hlova, I.Z.; Gupta, S.; Biswas, A.; Ma, T.; Zhou, L.; Balema, V.; Pecharsky, V.K.; Mudring, A.-V. Mechanochemical Synthesis, Luminescent and Magnetic Properties of Lanthanide Benzene-1,4-Dicarboxylate Coordination Polymers (Ln0.5Gd0.5)2 (1,4-BDC)3(H2O)4; Ln = Sm, Eu, Tb. New J. Chem. 2020, 44, 1054–1062. [Google Scholar] [CrossRef]
  13. Cabral Campello, M.P.; Palma, E.; Correia, I.; Paulo, P.M.R.; Matos, A.; Rino, J.; Coimbra, J.; Pessoa, J.C.; Gambino, D.; Paulo, A.; et al. Lanthanide Complexes with Phenanthroline-Based Ligands: Insights into Cell Death Mechanisms Obtained by Microscopy Techniques. Dalton Trans. 2019, 48, 4611–4624. [Google Scholar] [CrossRef] [PubMed]
  14. Decadt, R.; Van Hecke, K.; Depla, D.; Leus, K.; Weinberger, D.; Van Driessche, I.; Van Der Voort, P.; Van Deun, R. Synthesis, Crystal Structures, and Luminescence Properties of Carboxylate Based Rare-Earth Coordination Polymers. Inorg. Chem. 2012, 51, 11623–11634. [Google Scholar] [CrossRef] [PubMed]
  15. Massi, M.; Ogden, M. Luminescent Lanthanoid Calixarene Complexes and Materials. Materials 2017, 10, 1369. [Google Scholar] [CrossRef]
  16. Alpha, B.; Ballardini, R.; Balzani, V.; Lehn, J.; Perathoner, S.; Sabbatini, N. Antenna Effect in Luminescent Lanthanide Cryptates: A Photophysical Study. Photochem. Photobiol. 1990, 52, 299–306. [Google Scholar] [CrossRef]
  17. Orlova, A.V.; Kozhevnikova, V.Y.; Lepnev, L.S.; Goloveshkin, A.S.; Le-Deigen, I.M.; Utochnikova, V.V. NIR Emitting Terephthalates (SmxDyyGd1−xy)2(Tph)3(H2O)4 for Luminescence Thermometry in the Physiological Range. J. Rare Earths 2020, 38, 492–497. [Google Scholar] [CrossRef]
  18. Utochnikova, V.V.; Grishko, A.Y.; Koshelev, D.S.; Averin, A.A.; Lepnev, L.S.; Kuzmina, N.P. Lanthanide Heterometallic Terephthalates: Concentration Quenching and the Principles of the “Multiphotonic Emission”. Opt. Mater. 2017, 74, 201–208. [Google Scholar] [CrossRef]
  19. Nosov, V.G.; Toikka, Y.N.; Petrova, A.S.; Butorlin, O.S.; Kolesnikov, I.E.; Orlov, S.N.; Ryazantsev, M.N.; Kolesnik, S.S.; Bogachev, N.A.; Skripkin, M.Y.; et al. Brightly Luminescent (TbxLu1−x)2bdc3·nH2O MOFs: Effect of Synthesis Conditions on Structure and Luminescent Properties. Molecules 2023, 28, 2378. [Google Scholar] [CrossRef]
  20. Nosov, V.G.; Kupryakov, A.S.; Kolesnikov, I.E.; Vidyakina, A.A.; Tumkin, I.I.; Kolesnik, S.S.; Ryazantsev, M.N.; Bogachev, N.A.; Skripkin, M.Y.; Mereshchenko, A.S. Heterometallic Europium(III)–Lutetium(III) Terephthalates as Bright Luminescent Antenna MOFs. Molecules 2022, 27, 5763. [Google Scholar] [CrossRef]
  21. Kudyakova, Y.S.; Slepukhin, P.A.; Valova, M.S.; Burgart, Y.V.; Saloutin, V.I.; Bazhin, D.N. The Impact of the Alkali Metal Ion on the Crystal Structure and (Mechano)Luminescence of Terbium(III) Tetrakis(Β-diketonates). Eur. J. Inorg. Chem. 2020, 2020, 523–531. [Google Scholar] [CrossRef]
  22. Mironova, O.A.; Ryadun, A.A.; Sukhikh, T.S.; Konchenko, S.N.; Pushkarevsky, N.A. Synthesis and Luminescence Studies of Lanthanide Complexes (Gd, Tb, Dy) with Phenyl- and 2-Pyridylthiolates Supported by a Bulky β-Diketiminate Ligand. Impact of the Ligand Environment on Terbium(III) Emission. New J. Chem. 2020, 44, 19769–19779. [Google Scholar] [CrossRef]
  23. Judd, B.R. Optical Absorption Intensities of Rare-Earth Ions. Phys. Rev. 1962, 127, 750–761. [Google Scholar] [CrossRef]
  24. Ofelt, G.S. Intensities of Crystal Spectra of Rare-Earth Ions. J Chem Phys 1962, 37, 511–520. [Google Scholar] [CrossRef]
  25. Reineke, T.M.; Eddaoudi, M.; Fehr, M.; Kelley, D.; Yaghi, O.M. From Condensed Lanthanide Coordination Solids to Microporous Frameworks Having Accessible Metal Sites. J. Am. Chem. Soc. 1999, 121, 1651–1657. [Google Scholar] [CrossRef]
  26. Daiguebonne, C.; Kerbellec, N.; Guillou, O.; Bünzli, J.-C.; Gumy, F.; Catala, L.; Mallah, T.; Audebrand, N.; Gérault, Y.; Bernot, K.; et al. Structural and Luminescent Properties of Micro- and Nanosized Particles of Lanthanide Terephthalate Coordination Polymers. Inorg. Chem. 2008, 47, 3700–3708. [Google Scholar] [CrossRef]
  27. Feng, S. Poly[Hexaaquabis(μ3-Terephthalato)(μ2-Terephthalato)Diytterbium(III)]. Acta Crystallogr. Sect. E Struct. Rep. Online 2010, 66, m33. [Google Scholar] [CrossRef]
  28. Wang, P.; Li, Z.-F.; Song, L.-P.; Wang, C.-X.; Chen, Y. Catena -Poly[[[μ-Benzene-1,4-Dicarboxylato-Bis[Tetraaqualutetium(III)]]-Di-μ-Benzene-1,4-Dicarboxylato] Dihydrate]. Acta Crystallogr. Sect. E Struct. Rep. Online 2006, 62, m253–m255. [Google Scholar] [CrossRef]
  29. Mowat, J.P.S.; Miller, S.R.; Griffin, J.M.; Seymour, V.R.; Ashbrook, S.E.; Thompson, S.P.; Fairen-Jimenez, D.; Banu, A.-M.; Düren, T.; Wright, P.A. Structural Chemistry, Monoclinic-to-Orthorhombic Phase Transition, and CO2 Adsorption Behavior of the Small Pore Scandium Terephthalate, Sc2(O2CC6H4CO2)3, and Its Nitro- And Amino-Functionalized Derivatives. Inorg. Chem. 2011, 50, 10844–10858. [Google Scholar] [CrossRef]
  30. Pan, L.; Zheng, N.; Wu, Y.; Han, S.; Yang, R.; Huang, X.; Li, J. Synthesis, Characterization and Structural Transformation of A Condensed Rare Earth Metal Coordination Polymer. Inorg. Chem. 2001, 40, 828–830. [Google Scholar] [CrossRef]
  31. Holland, T.J.B.; Redfern, S.A.T. Unit Cell Refinement from Powder Diffraction Data: The Use of Regression Diagnostics. Mineral. Mag. 1997, 61, 65–77. [Google Scholar] [CrossRef]
  32. Pawley, G.S. Unit-Cell Refinement from Powder Diffraction Scans. J. Appl. Crystallogr. 1981, 14, 357–361. [Google Scholar] [CrossRef]
  33. Shannon, R.D. Revised Effective Ionic Radii and Systematic Studies of Interatomic Distances in Halides and Chalcogenides. Acta. Crystallogr. Sect. A 1976, 32, 751–767. [Google Scholar] [CrossRef]
  34. Bangham, A.D. A Correlation between Surface Charge and Coagulant Action of Phospholipids. Nature 1961, 192, 1197–1198. [Google Scholar] [CrossRef]
  35. Kolesnikov, I.E.; Povolotskiy, A.V.; Mamonova, D.V.; Kolesnikov, E.Y.; Kurochkin, A.V.; Lähderanta, E.; Mikhailov, M.D. Asymmetry Ratio as a Parameter of Eu3+ Local Environment in Phosphors. J. Rare Earths 2018, 36, 474–481. [Google Scholar] [CrossRef]
  36. Sudarsan, V.; van Veggel, F.C.J.M.; Herring, R.A.; Raudsepp, M. Surface Eu3+ Ions Are Different than “Bulk” Eu3+ Ions in Crystalline Doped LaF3 Nanoparticles. J. Mater. Chem. 2005, 15, 1332–1342. [Google Scholar] [CrossRef]
  37. Golyeva, E.V.; Kolesnikov, I.E.; Lähderanta, E.; Kurochkin, A.V.; Mikhailov, M.D. Effect of Synthesis Conditions on Structural, Morphological and Luminescence Properties of MgAl2O4:Eu3+ Nanopowders. J. Lumin. 2018, 194, 387–393. [Google Scholar] [CrossRef]
  38. Park, K.; Kim, H.; Hakeem, D.A. Effect of Host Composition and Eu3+ Concentration on the Photoluminescence of Aluminosilicate (Ca,Sr)2Al2SiO7:Eu3+ Phosphors. Dye. Pigment. 2017, 136, 70–77. [Google Scholar] [CrossRef]
  39. Oomen, E.W.J.L.; van Dongen, A.M.A. Europium (III) in Oxide Glasses. J. Non-Cryst. Solids 1989, 111, 205–213. [Google Scholar] [CrossRef]
  40. Schwarzenbach, G.; Flashka, H. Complexometric Titrations, 2nd ed.; Methuen: London, UK, 1969. [Google Scholar]
Figure 1. The PXRD patterns of (Eu1-xLax)2bdc3·nH2O (a) and (Eu1-xGdx)2bdc3·nH2O (b) (x = 0, 0.80, 0.90, 0.95, 0.98, 1) MOFs and the PXRD patterns of Eu2bdc3·4H2O [25] simulated from the single-crystals structures. The La3+ and Gd3+ contents are shown in the legends. The crystal structure of Eu2bdc3·4H2O adopted from ref. [20] (c).
Figure 1. The PXRD patterns of (Eu1-xLax)2bdc3·nH2O (a) and (Eu1-xGdx)2bdc3·nH2O (b) (x = 0, 0.80, 0.90, 0.95, 0.98, 1) MOFs and the PXRD patterns of Eu2bdc3·4H2O [25] simulated from the single-crystals structures. The La3+ and Gd3+ contents are shown in the legends. The crystal structure of Eu2bdc3·4H2O adopted from ref. [20] (c).
Molecules 29 00532 g001
Figure 2. The PXRD patterns of (Eu1-xLux)2bdc3·nH2O (x = 0, 0.80, 0.90, 0.95, 0.98, 1) MOFs and the PXRD patterns of Tb2bdc3 [25], Eu2bdc3·4H2O [25,26], Yb2bdc3·6H2O [27], Lu2bdc3·10H2O [28] simulated from the single-crystals structures. The Lu3+ content is shown in the legend.
Figure 2. The PXRD patterns of (Eu1-xLux)2bdc3·nH2O (x = 0, 0.80, 0.90, 0.95, 0.98, 1) MOFs and the PXRD patterns of Tb2bdc3 [25], Eu2bdc3·4H2O [25,26], Yb2bdc3·6H2O [27], Lu2bdc3·10H2O [28] simulated from the single-crystals structures. The Lu3+ content is shown in the legend.
Molecules 29 00532 g002
Figure 3. The PXRD patterns of calcinated (Eu1-xLux)2bdc3·nH2O (x = 0.95, 0.98, 1) compounds and the PXRD pattern of Tb2bdc3 [25] simulated from the single-crystals structure. The Lu3+ content is shown in the legend.
Figure 3. The PXRD patterns of calcinated (Eu1-xLux)2bdc3·nH2O (x = 0.95, 0.98, 1) compounds and the PXRD pattern of Tb2bdc3 [25] simulated from the single-crystals structure. The Lu3+ content is shown in the legend.
Molecules 29 00532 g003
Figure 4. Thermogravimetric curves of Lu2bdc3·nH2O (different line colors correspond to the three parallel measurements) (a) and Eu2bdc3·4H2O (b).
Figure 4. Thermogravimetric curves of Lu2bdc3·nH2O (different line colors correspond to the three parallel measurements) (a) and Eu2bdc3·4H2O (b).
Molecules 29 00532 g004
Figure 5. SEM images of the samples (Eu1-xLnx)2bdc3·nH2O (Ln = Lu, La, Gd).
Figure 5. SEM images of the samples (Eu1-xLnx)2bdc3·nH2O (Ln = Lu, La, Gd).
Molecules 29 00532 g005
Figure 6. Emission spectra of (Eu1-xGdx)2bdc3·nH2O (a) and (Eu1-xLax)2bdc3·nH2O (b) (x = 0, 0.80, 0.90, 0.95, 0.98) compounds upon 310-nm excitation. The La3+ and Gd3+ contents are shown in the legends.
Figure 6. Emission spectra of (Eu1-xGdx)2bdc3·nH2O (a) and (Eu1-xLax)2bdc3·nH2O (b) (x = 0, 0.80, 0.90, 0.95, 0.98) compounds upon 310-nm excitation. The La3+ and Gd3+ contents are shown in the legends.
Molecules 29 00532 g006
Figure 7. Emission spectra of the (Eu1-xLux)2bdc3·nH2O nH2O (x = 0, 0.80, 0.90, 0.95, 0.98) compounds upon 310-nm excitation. The Lu3+ content is shown in the legend.
Figure 7. Emission spectra of the (Eu1-xLux)2bdc3·nH2O nH2O (x = 0, 0.80, 0.90, 0.95, 0.98) compounds upon 310-nm excitation. The Lu3+ content is shown in the legend.
Molecules 29 00532 g007
Figure 8. The photoluminescence decay curves of the (Eu1-xGdx)2bdc3·nH2O (a) and (Eu1-xLax)2bdc3·nH2O (b) (x = 0, 0.80, 0.90, 0.95, 0.98) compounds monitored at 614 nm upon 310-nm excitation.
Figure 8. The photoluminescence decay curves of the (Eu1-xGdx)2bdc3·nH2O (a) and (Eu1-xLax)2bdc3·nH2O (b) (x = 0, 0.80, 0.90, 0.95, 0.98) compounds monitored at 614 nm upon 310-nm excitation.
Molecules 29 00532 g008
Figure 9. The photoluminescence decay curves of (Eu1-xLux)2bdc3·nH2O nH2O (x = 0, 0.80, 0.90, 0.95, 0.98) were monitored at 614 nm.
Figure 9. The photoluminescence decay curves of (Eu1-xLux)2bdc3·nH2O nH2O (x = 0, 0.80, 0.90, 0.95, 0.98) were monitored at 614 nm.
Molecules 29 00532 g009
Table 1. Unit cell parameters with calculation errors for (Eu1-xLnx)2bdc3·nH2O (Ln = Lu, La, Gd) refined for Tb2bdc3·4H2O crystalline phase.
Table 1. Unit cell parameters with calculation errors for (Eu1-xLnx)2bdc3·nH2O (Ln = Lu, La, Gd) refined for Tb2bdc3·4H2O crystalline phase.
χLu (at.%)a, Åb, Åc, ÅαβγV, Å3
06.1860 ± 0.001810.103 ± 0.00310.184 ± 0.003102.279 ± 0.02691.423 ± 0.027101.482 ± 0.028608.00 ± 0.24
806.1435 ± 0.001810.014 ± 0.00310.061 ± 0.003101.985 ± 0.02691.683 ± 0.026101.232 ± 0.028592.26 ± 0.23
906.0896 ± 0.00189.994 ± 0.0039.971 ± 0.003101.466 ± 0.02691.631 ± 0.026100.876 ± 0.028582.66 ± 0.23
χLa (at.%)a, Åb, Åc, ÅαβγV, Å3
06.1860 ± 0.001810.103 ± 0.00310.184 ± 0.003102.279 ± 0.02691.423 ± 0.027101.482 ± 0.028608.00 ± 0.24
806.2261 ± 0.001810.142 ± 0.00310.273 ± 0.003102.127 ± 0.02691.657 ± 0.027101.507 ± 0.028619.84 ± 0.24
906.2477 ± 0.001910.198 ± 0.00410.292 ± 0.003102.169 ± 0.02691.599 ± 0.027101.513 ± 0.028626.41 ± 0.25
956.2632 ± 0.001910.197 ± 0.00410.308 ± 0.003102.235 ± 0.02691.537 ± 0.027101.565 ± 0.028628.68 ± 0.25
986.2412 ± 0.000510.159 ± 0.00810.283 ± 0.013102.32± 0.1891.53 ± 0.15101.78 ± 0.12621.85 ± 0.25
1006.2691 ± 0.001910.206 ± 0.00410.336 ± 0.003102.260 ± 0.02691.552 ± 0.027101.495 ± 0.028631.61 ± 0.25
χGd (at.%)a, Åb, Åc, ÅαβγV, Å3
06.1860 ± 0.001810.103 ± 0.00310.184 ± 0.003102.279 ± 0.02691.423 ± 0.027101.482 ± 0.028608.00 ± 0.24
806.2368 ± 0.001910.036 ± 0.00310.251 ± 0.003102.879 ± 0.02691.989 ± 0.027101.617 ± 0.028610.54 ± 0.23
906.1611 ± 0.001810.077 ± 0.00310.148 ± 0.003102.212 ± 0.02691.258 ± 0.026101.411 ± 0.028602.25 ± 0.23
956.2354 ± 0.001910.031 ± 0.00310.234 ± 0.003102.829 ± 0.02691.931 ± 0.027101.480 ± 0.028609.61 ± 0.23
986.2235 ± 0.00189.995 ± 0.00310.238 ± 0.003103.065 ± 0.02691.812 ± 0.027101.641 ± 0.027605.60 ± 0.23
1006.2112 ± 0.00189.992 ± 0.00310.227 ± 0.003102.810 ± 0.02692.048 ± 0.027101.481 ± 0.027604.40 ± 0.23
Table 2. Sizes of particles of the samples (Eu1-xLax)2bdc3·nH2O, (Eu1-xGdx)2bdc3·nH2O, and (Eu1-xLux)2bdc3·nH2O.
Table 2. Sizes of particles of the samples (Eu1-xLax)2bdc3·nH2O, (Eu1-xGdx)2bdc3·nH2O, and (Eu1-xLux)2bdc3·nH2O.
(Eu1-xLax)2bdc3·nH2O(Eu1-xGdx)2bdc3·nH2O(Eu1-xLux)2bdc3·nH2O
χLa (at.%)width, μmlength, μmχGd (at.%)width, μmlength, μmχLu (at.%)width, μmlength, μm
1002.7 ± 1.28.7 ± 2.21002.2 ± 1.35.8 ± 2.71008.2 ± 3.612 ± 4
981.9 ± 0.65.3 ± 1.4982.7 ± 1.15.8 ± 1.9985.1 ± 1.58.0 ± 2.3
953.1 ± 0.98.5 ± 2.5951.7 ± 0.64.2 ± 1.1954.6 ± 1.27.5 ± 1.2
902.8 ± 1.07.4 ± 2.1903.0 ± 0.87.4 ± 2.1900.8 ± 0.34.9 ± 1.3
802.7 ± 1.06.5 ± 1.7803.0 ± 1.07.0 ± 2.1800.8 ± 0.43.5 ± 1.6
01.6 ± 0.83.9 ± 1.901.6 ± 0.83.9 ± 1.901.6 ± 0.83.9 ± 1.9
Table 3. Lifetimes (τ) and photoluminescence quantum yields (ΦPL) of (Eu1-xLnx)2bdc3·nH2O (Ln = La, Gd, Lu; x = 0, 0.80, 0.90, 0.95, 0.98). The fractions of the exponential components are given in parentheses.
Table 3. Lifetimes (τ) and photoluminescence quantum yields (ΦPL) of (Eu1-xLnx)2bdc3·nH2O (Ln = La, Gd, Lu; x = 0, 0.80, 0.90, 0.95, 0.98). The fractions of the exponential components are given in parentheses.
(Eu1-xLax)2bdc3·nH2O(Eu1-xGdx)2bdc3·nH2O(Eu1-xLux)2bdc3·nH2O
χLn (at.%)τ, msΦPL, %τ, msΦPL, %τ1, msτ2, msΦPL, %
980.45 ± 0.019 ± 10.47 ± 0.0110 ± 10.41 ± 0.01 (91.4%)1.54 ± 0.18 (8.6%)16 ± 1
950.44 ± 0.0110 ± 10.46 ± 0.0112 ± 10.34 ± 0.01 (85.9%)1.62 ± 0.08 (14.1%)23 ± 1
900.45 ± 0.019 ± 10.46 ± 0.0113 ± 10.46 ± 0.01-11 ± 1
800.44 ± 0.0110 ± 10.46 ± 0.0112 ± 10.46 ± 0.01-12 ± 1
00.43 ± 0.0110 ± 10.43 ± 0.0110 ± 10.43 ± 0.01-10 ± 1
Table 4. Asymmetry ratio Ras of (Eu1-xLnx)2bdc3·nH2O (Ln = La, Gd, Lu; x = 0, 0.80, 0.90, 0.95, 0.98).
Table 4. Asymmetry ratio Ras of (Eu1-xLnx)2bdc3·nH2O (Ln = La, Gd, Lu; x = 0, 0.80, 0.90, 0.95, 0.98).
χLa (at.%)(Eu1-xLax)2bdc3·nH2O(Eu1-xGdx)2bdc3·nH2O(Eu1-xLux)2bdc3·nH2O
983.8 ± 0.33.8 ± 0.33.0 ± 0.3
953.9 ± 0.33.8 ± 0.32.9 ± 0.3
903.9 ± 0.33.9 ± 0.33.5 ± 0.3
803.9 ± 0.33.7 ± 0.33.7 ± 0.3
03.7 ± 0.33.7 ± 0.33.7 ± 0.3
Table 5. The concentrations of REE ions in initial solutions used for the synthesis of (Eu1-xLnx)2bdc3·nH2O MOFs.
Table 5. The concentrations of REE ions in initial solutions used for the synthesis of (Eu1-xLnx)2bdc3·nH2O MOFs.
χLn (at.%)C(Eu3+), mMC(Ln3+), mM
10005.0
980.14.9
950.254.75
900.54.5
801.04.0
05.00
Table 6. Eu3+ relative atomic fraction to the total amount of Eu3+ and Ln3+ (Ln = Lu, La, Gd) in (Eu1-xLnx)2bdc3·nH2O compounds (x = 0, 0.80, 0.90, 0.95, 0.98, 1) taken during synthesis and obtained from EDX data.
Table 6. Eu3+ relative atomic fraction to the total amount of Eu3+ and Ln3+ (Ln = Lu, La, Gd) in (Eu1-xLnx)2bdc3·nH2O compounds (x = 0, 0.80, 0.90, 0.95, 0.98, 1) taken during synthesis and obtained from EDX data.
(Eu1-xLax)2bdc3·nH2O(Eu1-xGdx)2bdc3·nH2O(Eu1-xLux)2bdc3·nH2O
χLa (at.%), TakenχLa (at.%), EDXχGd (at.%), TakenχGd (at.%), EDXχLu (at.%), TakenχLu (at.%), EDX
100100100100100100
9897.4 ± 0.99898 ± 19898 ± 1
9594.7 ± 0.69595.4 ± 0.69594 ± 3
9089.4 ± 0.59090.2 ± 0.89091 ± 4
8079.7 ± 2.38081.1 ± 1.28079 ± 6
000000
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Kolesnik, S.S.; Bogachev, N.A.; Kolesnikov, I.E.; Orlov, S.N.; Ryazantsev, M.N.; González, G.; Skripkin, M.Y.; Mereshchenko, A.S. Microcrystalline Luminescent (Eu1-xLnx)2bdc3·nH2O (Ln = La, Gd, Lu) Antenna MOFs: Effect of Dopant Content on Structure, Particle Morphology, and Luminescent Properties. Molecules 2024, 29, 532. https://doi.org/10.3390/molecules29020532

AMA Style

Kolesnik SS, Bogachev NA, Kolesnikov IE, Orlov SN, Ryazantsev MN, González G, Skripkin MY, Mereshchenko AS. Microcrystalline Luminescent (Eu1-xLnx)2bdc3·nH2O (Ln = La, Gd, Lu) Antenna MOFs: Effect of Dopant Content on Structure, Particle Morphology, and Luminescent Properties. Molecules. 2024; 29(2):532. https://doi.org/10.3390/molecules29020532

Chicago/Turabian Style

Kolesnik, Stefaniia S., Nikita A. Bogachev, Ilya E. Kolesnikov, Sergey N. Orlov, Mikhail N. Ryazantsev, Gema González, Mikhail Yu. Skripkin, and Andrey S. Mereshchenko. 2024. "Microcrystalline Luminescent (Eu1-xLnx)2bdc3·nH2O (Ln = La, Gd, Lu) Antenna MOFs: Effect of Dopant Content on Structure, Particle Morphology, and Luminescent Properties" Molecules 29, no. 2: 532. https://doi.org/10.3390/molecules29020532

Article Metrics

Back to TopTop