Next Article in Journal
Preparation and Surface Characterization of Chitosan-Based Coatings for PET Materials
Next Article in Special Issue
Computation of Entropy Measures for Metal-Organic Frameworks
Previous Article in Journal
Different Types of Hypericum perforatum cvs. (Elixir, Helos, Topas) In Vitro Cultures: A Rich Source of Bioactive Metabolites and Biological Activities of Biomass Extracts
Previous Article in Special Issue
Dye-Encapsulated Metal–Organic Frameworks for the Multi-Parameter Detection of Temperature
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Brightly Luminescent (TbxLu1−x)2bdc3·nH2O MOFs: Effect of Synthesis Conditions on Structure and Luminescent Properties

by
Viktor G. Nosov
1,
Yulia N. Toikka
1,
Anna S. Petrova
1,
Oleg S. Butorlin
1,
Ilya E. Kolesnikov
1,
Sergey N. Orlov
1,2,3,
Mikhail N. Ryazantsev
1,4,
Stefaniia S. Kolesnik
1,
Nikita A. Bogachev
1,
Mikhail Yu. Skripkin
1 and
Andrey S. Mereshchenko
1,*
1
Saint-Petersburg State University, 7/9 Universitetskaya emb., 199034 St. Petersburg, Russia
2
Federal State Unitary Enterprise “Alexandrov Research Institute of Technology”, 72 Koporskoe Shosse, 188540 Sosnovy Bor, Russia
3
Institute of Nuclear Industry, Peter the Great St. Petersburg Polytechnic University (SPbSU), 29 Polytechnicheskaya Street, 195251 St. Petersburg, Russia
4
Nanotechnology Research and Education Centre RAS, Saint Petersburg Academic University, ul. Khlopina 8/3, 194021 St. Petersburg, Russia
*
Author to whom correspondence should be addressed.
Molecules 2023, 28(5), 2378; https://doi.org/10.3390/molecules28052378
Submission received: 27 February 2023 / Revised: 2 March 2023 / Accepted: 2 March 2023 / Published: 4 March 2023
(This article belongs to the Special Issue Multifunctional Metal-Organic Framework Materials)

Abstract

:
Luminescent, heterometallic terbium(III)–lutetium(III) terephthalate metal-organic frameworks (MOFs) were synthesized via direct reaction between aqueous solutions of disodium terephthalate and nitrates of corresponding lanthanides by using two methods: synthesis from diluted and concentrated solutions. For (TbxLu1−x)2bdc3·nH2O MOFs (bdc = 1,4-benzenedicarboxylate) containing more than 30 at. % of Tb3+, only one crystalline phase was formed: Ln2bdc3·4H2O. At lower Tb3+ concentrations, MOFs crystallized as the mixture of Ln2bdc3·4H2O and Ln2bdc3·10H2O (diluted solutions) or Ln2bdc3 (concentrated solutions). All synthesized samples that contained Tb3+ ions demonstrated bright green luminescence upon excitation into the 1ππ* excited state of terephthalate ions. The photoluminescence quantum yields (PLQY) of the compounds corresponding to the Ln2bdc3 crystalline phase were significantly larger than for Ln2bdc3·4H2O and Ln2bdc3·10H2O phases due to absence of quenching from water molecules possessing high-energy O-H vibrational modes. One of the synthesized materials, namely, (Tb0.1Lu0.9)2bdc3·1.4H2O, had one of the highest PLQY among Tb-based MOFs, 95%.

Graphical Abstract

1. Introduction

Rare earth elements (REE)-based compounds are promising materials for applications in medicine [1,2], sensors [3,4], catalysis [5], anticounterfeiting [6,7], bioimaging [8,9], photovoltaic systems [10,11,12], etc. due to their unique optical and magnetic properties. The positions of the narrow emission bands of REE ions attributed to f–f transitions strongly depend only on the type of lanthanide ions. This property allows photoluminescence color tuning of the REE-containing materials [13]. Usually, purely inorganic compounds of lanthanides demonstrate relatively weak photoluminescence intensity because they possess extremely low extinction coefficients due to the forbidden nature of f–f transitions, which makes the direct excitation of ions inefficient. This issue can be overcome using the so-called “antenna effect”. The antenna effect is realized in some metal–organic compounds in which the light is absorbed by the chromophore group of the organic ligand followed by the energy transfer to the lanthanide ion, which then emits the light corresponding to the characteristic f–f transitions [14,15,16]. The typical ligands used in REE antenna complexes are calixarenes [17], dipicolinic acid [18], tris-bipyridines [19], and carboxylates including terephthalates [20,21,22]. Lanthanide-based metal–organic frameworks (MOFs) combine the optical properties of REE-based materials with the topological features of MOFs, which makes them exceptional materials for chemical sensors [23], optical thermometers [24,25], and OLED components [26,27]. Eu3+ and Tb3+ ions are often used as activators in such materials because of their strong red and green emissions, respectively [21,28,29,30]. The simultaneous presence of several lanthanide ions in one compound provides the possibility to gain the properties of multimodal imaging agents and allows one to discover the energy transfer mechanisms in such compounds [31,32,33]. Moreover, some studies showed the enhancement of luminescence of Eu3+, Tb3+, and Sm3+ containing antenna MOFs upon dilution with paramagnetic Gd3+ ions, whereas the substitution of luminescent REE ions by diamagnetic La3+, Y3+, and Lu3+ ions does not lead to the luminescence intensity increase [20]. It is important to note that the doping by the aforementioned ions does not result in the crystalline phase change in the majority of studies of heterometallic REE MOFs. At the same time, the number of studies of luminescent antenna MOFs containing both luminescent and non-luminescent REE ions is still insignificant in contrast with those of solid solutions of purely inorganic compounds [34,35,36,37,38]. Recently, we studied the optical properties of heterometallic europium–lutetium terephthalates [39]. The luminescence quantum yields of the terephthalate ions were found to be increased with a decrease in the europium concentrations in these compounds. We also observed that the substitution of a large amount of Eu3+ for Lu3+ resulted in a crystalline phase change from Ln2bdc3·4H2O to Ln2bdc3 (bdc = 1,4-benzenedicarboxylate). The lifetimes of europium (III)’s 5D0 excited state were found to be larger by 4–4.8 times in an anhydrous phase with low Eu3+ content.
In order to further reveal the doping effect of Lu3+ ions on the structural and optical properties of antenna luminescent MOFs, in the current work, we studied bimetallic terbium(III)–lutetium(III) terephthalates as obtained by using two methods.

2. Results and Discussion

2.1. PXRD Results and Analysis

All of the syntheses, whatever the chosen method, yielded crystalline samples. In Figure 1a,b, the PXRD patterns of the (TbxLu1−x)2bdc3·nH2O (x = 0–1) MOFs synthesized from diluted and concentrated solutions are shown. We found that all compounds with concentration of terbium (III) ions 30 at. % and more were isostructural to the Ln2bdc3·4H2O crystalline phase (Ln = Ce − Yb) [40], and additional peaks were not observed. At low Tb3+ concentrations between 0 and 5 at. %, the positions of the reflexes in the PXRD patterns were different from those of the Ln2bdc3·4H2O and depended on the concentration of the initial reagents (Na2bdc, TbCl3, and LuCl3). Thus, diffraction patterns corresponded to Ln2bdc3·10H2O [41] and Ln2bdc3 [42] for compounds synthesized from diluted and concentrated solutions, respectively (see Section 3). At intermediate Tb3+ concentrations between 10 and 25 at. %, the binary mixtures of the aforementioned crystalline phases were precipitated, namely, Ln2bdc3·4H2O + Ln2bdc3·10H2O and Ln2bdc3·4H2O + Ln2bdc3 for the (TbxLu1−x)2bdc3·nH2O MOFs obtained from diluted and concentrated solutions, respectively.
The effect of molar ratio of reagents taken for synthesis on MOFs structure was observed previously (see, for example, [43,44]), but was not explained properly. It is generally accepted that MOFs are formed stepwise from the secondary building units (SBUs), metal–ligand oligomers that replicates themselves to form MOF-like structures [45]. Therefore, the final structure of coordination polymer allows us to assume the possible reasons behind the differences between compounds of two synthesized series. The crystal structures of Ln2bdc3·4H2O, Ln2bdc3, and Ln2bdc3·10H2O are shown in Figure 2. In Ln2bdc3·4H2O lanthanide (III), ions were bound to two water molecules and six terephthalate ions through oxygen atoms, where Ln3+ coordination number (CN) is equal to 8. In Ln2bdc3 structures, Ln3+ ions with CN = 7 coordinated solely to oxygens of terephthalate ions. In the Ln2bdc3·10H2O structure, the metal center coordination number was also equal to 7, but four coordination sites were occupied by water molecules. Two water molecules per one formula unit in the Ln2bdc3·10H2O structure were located in interplanar channels. Commonly, Tb3+ ions have relatively larger coordination numbers than Lu3+ ions. For example, in aqueous solutions, Tb3+ dominantly exists in a nonacoordinated form as the [Tb(H2O)9]3+ complex [46], but smaller Yb3+ and Lu3+ ions possess lower coordination numbers and exist as [Ln(H2O)8]3+ [47,48]. Therefore, we expected that terbium ions will reveal larger coordination numbers than lutetium ion in our MOFs. Indeed, the analysis of the aforementioned structures (Figure 2) revealed that lanthanide ions have coordination numbers of seven in Ln2bdc3 and Ln2bdc3·10H2O, which are formed in pure lutetium terephthalate, and in (TbxLu1−x)2bdc3·nH2O at high Lu3+ content levels. In Ln2bdc3·4H2O, which is formed in pure terbium terephthalate and in mixed Tb–Lu terephthalates at high Tb3+ content levels, the coordination number of Ln3+ ions is equal to eight. The reasons that can explain the difference between structures of lutetium terephthalates synthesized from diluted (Lu2bdc3·10H2O) and concentrated (Lu2bdc3) solutions are unclear. We assume the key factor that affects the structure of precipitated MOF is the fractional distribution of initially formed metastable complexes [Ln(H2O)x(bdc)y]3−2y [49]. These complexes then aggregate into SBUs, which further form MOFs. Apparently, in concentrated solutions, complexes have higher Lu3+:bdc2− ratios (1:2 or 1:3) than in diluted solution (1:1). Therefore, further formed SBUs and MOFs of Lu2bdc3 had larger number of coordinated oxygens of terephthalate ligands than Lu2bdc3·10H2O.
In our previous work, we reported the similar behavior of the (EuxLu1−x)2bdc3·nH2O MOFs obtained from concentrated solutions [39]. We found that phase transition occurred at significantly lower Eu3+ concentrations (6 at. % of Eu3+ vs. 30 at. % of Tb3+). This observation can be explained by the lower ionic radius of Tb3+ (1.040Å) than that of Eu3+ (1.066Å) [50]. The structure with CN = 7 (Ln2bdc3) is more advantageous for Tb3+ than for Eu3+, which forms a structure Ln2bdc3·4H2O with larger CN = 8 beginning at 6 at. % of Eu3+ ions.

2.2. Thermogravimetric Analysis (TGA)

The thermal behavior of the selected compounds (TbxLu1−x)2bdc3·nH2O (x = 0–1) was studied by using the thermogravimetric method (TGA). The TGA curves of the MOFs obtained from diluted and concentrated solutions were recorded in the temperature range of 35–200 °C (Figure 3). When heated, the lanthanide terephthalates decomposed in two common steps: (i) dehydration of the compounds, resulting in formation of Ln2bdc3 at about 100–200 °C, and (ii) the structural decomposition of coordination polymers [42]. The observed weight loss at 100–190 °C for all measured samples corresponded to the dehydration step; therefore, the analysis of the TGA curves allowed us to calculate the average numbers of water molecules in the coordination polymers (TbxLu1−x)2bdc3·nH2O.
The number of water molecules per one formula unit N(H2O) for all selected compounds as function of Tb3+ concentration is shown in Figure 4a,b for samples synthesized from diluted and concentrated solutions, respectively. The number of water molecules per one formula unit is equal to four for pure terbium terephthalate (100 at. % Tb3+) in both series. The N(H2O) value increases from 4 to 10 upon the substitution of Tb3+ by Lu3+ ions (decreasing of Tb3+ content) in (TbxLu1−x)2bdc3·nH2O MOFs obtained from the diluted solutions (Figure 4a). However, for MOFs synthesized from the diluted solutions, the number of water molecules decreases from four to zero upon Tb3+ concentration decrease (Figure 4b). These facts are in agreement with the XRD data, in which we observed phase transitions from Ln2bdc3·4H2O either to Ln2bdc3·10H2O or Ln2bdc3 upon the decrease of Tb3+ content. Summarizing the TGA and XRD data, we estimated the molar fraction of each coexisting crystalline phase (Figure 4c,d). The molar fraction of Ln2bdc3·4H2O increased from 0 to 30 at. % Tb3+ for two synthesized series of MOFs (TbxLu1−x)2bdc3·nH2O, and simultaneously, the molar fraction of the second coexisting phase decreased. In the Tb3+ concentration range of 30–100 at. %, only Ln2bdc3·4H2O was present in both series.

2.3. Luminescent Properties

Aromatic carboxylate ions, especially benzene dicarboxylates, are typical linkers for the luminescent antenna MOF design [15,20] due to the efficient sensitization of lanthanide luminescence. The sensitization mechanism consists of several steps. Upon UV-photon absorption, the linker is promoted into the Sn(1ππ*) exited electronic state, which is followed by the fast internal conversion to S1(1ππ*). Due to the heavy atom effect, the S1 state of the linker efficiently undergoes intersystem crossing to the T1(3ππ*) triplet electronic excited state [32]. If the T1 state of organic linker lies slightly higher in energy than one of the levels of activator lanthanide ion, then the energy is efficiently transferred to the lanthanide ion and followed by the photon emission corresponding to the f–f transition. Thus, terbium terephthalate, Tb2bdc3·4H2O, demonstrates a relatively high Tb3+ photoluminescence quantum yield (43–55% [32,42,51]) upon UV-excitation into terephthalate ions due to the fact that the T1 state of the terephthalate ion (E(T1) ≈ 20,400–20,650 cm−1 [32] for bdc2-) lies only 50—300 cm−1 above the 5D4 level of the Tb3+ ion (E(5D4) ≈ 20,350 cm−1 [52]).
The emission spectra of the synthesized compounds, which were measured upon 280-nm excitation into the Sn(1ππ*) excited electronic state of terephthalate ions, are shown in Figure 5. The observed emission spectra are typical for compounds containing Tb3+ ions [53] and consist of narrow bands corresponding to 5D47FJ (J = 3–6) transitions of Tb3+: 5D47F6 (≈491 nm), 5D47F5 (≈543 nm), 5D47F4 (≈585 nm), and 5D47F3 (≈620 nm). One can observe that the fine structure of Tb3+ emission spectra of (TbxLu1−x)2bdc3·nH2O significantly changes at Tb3+ concentration of about 20 at. % in both studied series. It is well-known, that the fine structure of lanthanide (III) ions strictly depends on the local symmetry of emitting lanthanide ions [54,55,56,57]. Indeed, one can notice three different types of fine structure of the spectra: (i) compounds with terbium (III) content of 25 at. % and more in both series (corresponding to the (TbxLu1−x)2bdc3·4H2O structure that dominates in this range of concentrations); (ii) MOFs with Tb3+ concentrations less than 25 at. % in series obtained from diluted solutions ((TbxLu1−x)2bdc3·10H2O as the dominating structure); (iii) compounds with terbium (III) concentrations less than 25 at. % in series obtained from concentrated solutions ((TbxLu1−x)2bdc3 as the dominating structure). The difference is that the fine structure of the emission bands is attributed to the different symmetry of the first coordination sphere of the Tb3+ ion in these three types of crystalline structures.
Figure 6 displays the photoluminescence decay curves measured upon UV-excitation of (TbxLu1−x)2bdc3·nH2O MOFs synthesized via the two methods mentioned as monitored at 543 nm (5D47F5 transition). At terbium (III) ion concentrations of 60 and 100 at. %, photoluminescence decay curves were well-fitted with the single exponential functions (Equation (1)) with time constants τ of about 0.7–1.1 ms. At low levels of Tb3+ content (1, 5, and 10 at. %), the photoluminescence decay curves of the compounds obtained from concentrated solutions fit the double exponential functions (Equation (2)). The biexponential behavior of the photoluminescence decay indicates the presence of different relaxation pathways of Tb3+ ions corresponding to two terbium ions with different coordination environments. We believe that the larger time constant τ2, which is about 2.6–3 ms (Table 1), corresponds to lifetime of 5D4 state Tb3+ ions in the (TbxLu1−x)2bdc3 structure. The smaller time constant τ1 value (1.0–1.5 ms) can be assigned to the lifetime of the 5D4 state of terbium (III) ions in the (TbxLu1−x)2bdc3·4H2O structure. The photoluminescence decay curves of the (TbxLu1−x)2bdc3·nH2O compounds with x = 0.01, 0.05, and 0.10, which were obtained from diluted solutions, fit the single exponential functions (eq. 1) with time constants of about 1.1 ms. As the XRD and TGA data shows the coexistence of Ln2bdc3·4H2O and Ln2bdc3·10H2O phases in these compounds, one would expect the presence of two different exponential components of photoluminescence decay curves. Most likely, the values of the 5D4 energy level lifetime of Tb3+ ions in Ln2bdc3·4H2O and Ln2bdc3·10H2O structures are close to each other, as pseudo-single-exponential decay was observed.
I = I 1 · e t τ
I = I 1 · e t τ 1 + I 2 · e t τ 2
We have found that the 5D4 excited state lifetimes in the (TbxLu1−x)2bdc3·nH2O MOFs obtained from diluted solutions decreased from 1.122 to 0.696 ms with the increase of terbium concentration due to the increased probability of energy transfer between neighboring Tb3+ ions with subsequent quenching of impurities. At the same time, the photoluminescent quantum yields (PLQY) of these compounds had maxima at about 60 at. % of Tb3+, where PLQY is equal to 60% (Table 1). Typically, emission intensity and PLQY nonlinearly depend on the concentration of Tb3+ ions [58,59]. This type of concentration dependence can be explained by the two competitive effects in REE-containing phosphors [60,61]. Thus, the rise of the numbers of luminescent sites results in radiative emission probability increased and, as a result, the emission intensity and PLQY increased. At the same time, upon the Tb3+ concentration’s rise, the distance between Tb3+ ions decreased, resulting in the nonradiative processes probability increase that led to the emission quenching [62], resulting in lower PLQY values of pure terbium terephthalate (100 at.% of Tb3+) relative to the MOFs containing 60 at.% of Tb3+. The PLQY of the (TbxLu1−x)2bdc3·nH2O MOFs obtained from concentrated solutions are equal to the ones obtained from the diluted solutions at the Tb3+ concentration of 60 and 100 at. %, where the MOFs formed in the same crystalline phase, namely, Ln2bdc3·4H2O. A further decrease of Tb3+ content in the MOFs obtained from the diluted solutions resulted in a significant PLQY rise, reaching maxima of 95% for the (Tb0.1Lu0.9)2bdc3·1.4H2O sample. The higher values of PLQY and excited state lifetimes of these materials are attributed to the formation of the anhydrous Ln2bdc3 crystalline phase. The PLQY of the Ln2bdc3 MOFs were significantly higher than that of the Ln2bdc3·4H2O and Ln2bdc3·10H2O MOFs due to the absence of water molecules coordinated to Tb3+ ions, which efficiently quenches luminescence due to energy transfer from the 5D4 excited state of Tb3+ ions to the high-energy O-H stretching vibrational modes of H2O molecules [63].

3. Materials and Methods

Benzene-1,4-dicarboxylic (terephthalic, H2bdc) acid (>98%), sodium hydroxide (>99%), nickel(II) chloride hexahydrate (>99%), EDTA disodium salt (0.05M aqueous solution), and murexide were purchased from Sigma-Aldrich Chemie GmbH (Taufkirchen, Germany) and used without additional purification. Lutetium (III) nitrate pentahydrate and terbium (III) nitrate pentahydrate were purchased from Chemcraft (Kaliningrad, Russia). The 0.3 M solution of disodium terephthalate (Na2bdc) was prepared by dissolving 0.6 moles of sodium hydroxide and 0.3 moles of terephthalic acid in 1 L of distilled water. Volumes of 0.2 M of TbCl3 and LuCl3 solutions were prepared and standardized using back complexometric titration. Thus, 1 mL of LnCl3 (Ln = Tb, Lu) solution with a concentration of about 0.3 M, 20 mL of 0.05 M EDTA, 10 mL of ammonium buffer solution (pH = 9), and a pinch of murexide indicator were taken in a conical flask. The obtained solution was titrated with 0.05 M NiCl2 [64]. Then, standardized LnCl3 solutions were diluted to 0.2 M.
White powders of the (TbxLu1−x)2bdc3·nH2O MOFs were synthesized by the direct mixing of two aqueous solutions: (1) sodium terephthalate and (2) terbium and lutetium nitrates taken in various ratios, as shown in Table 2. In order to reveal the effect of the concentrations of the initial solutions on the properties of the obtained materials, we synthesized two series of (TbxLu1−x)2bdc3·nH2O MOFs. Series 1 was obtained from the Na2bdc and LnCl3 diluted solutions, where 8 mL of 0.1 M Na2bdc solution was added dropwise under vigorous stirring to a solution containing 5 mL of distilled water and 2 mL of 0.2M TbCl3 and LuCl3 solutions taken in certain ratios (Table 2). Series 2 was obtained from the Na2bdc and LnCl3 concentrated solutions, where 3 mL 0.3 M Na2bdc solution was rapidly added to the 2 mL of TbCl3 and LuCl3 solutions taken in various ratios, as shown in Table 2. Obtained suspensions were kept for one 1 h at room temperature, and then, solid precipitates of the (TbxLu1−x)2bdc3·nH2O MOFs were separated from the reaction mixture via centrifugation (2300 g) and washed with deionized water 5 times. The resulting white powders of terbium-lutetium terephthalates were dried in an air atmosphere at 60 °C for 24 h.
The Tb3+/Lu3+ ratios in the synthesized (TbxLu1−x)2bdc3·nH2O compounds were confirmed with energy-dispersive X-ray spectroscopy (EDX) (EDX spectrometer EDX-800P, Shimadzu, Japan) (Table 3). We found that the amounts of the elements are consistent with experimental EDX data. The X-ray powder diffraction (XRD) data of obtained (TbxLu1−x)2bdc3·nH2O samples were taken with a D2 Phaser (Bruker, Billerica, MA, USA) X-ray diffractometer using Cu Kα radiation (λ = 1.54056 Å). The thermal behavior of the compounds was studied via 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). Solid-state luminescence emission spectra were recorded with a Fluoromax-4 fluorescence spectrometer (Horiba Jobin–Yvon, Kyoto, Japan). Lifetime measurements were performed with the same spectrometer using a pulsed Xe lamp (pulse duration 3 µs). The quantum yield measurements were performed by using the Fluorolog 3 Quanta-phi device (Horiba Jobin–Yvon, Kyoto, Japan).

4. Conclusions

In this work, we reported the phase composition and the optical properties of luminescent antenna MOFs: heterometallic terbium(III)–lutetium(III) terephthalates. The series of (TbxLu1−x)2bdc3·nH2O (x = 0–1) were synthesized via direct reaction between aqueous solutions of disodium terephthalate and nitrates of corresponding lanthanides with two methods: using diluted and concentrated solutions. At Tb3+ concentrations more than 25 at. %, synthesized compounds existed in the Ln2bdc3·4H2O crystal structure with the coordination number (CN) of the lanthanide ion equal to eight. Lu3+ ions typically have lower coordination numbers than Tb3+ ions; hence, at high lutetium (III) content, structures with CN(Ln3+) < 8 crystallized. Therefore, compounds containing small amounts of terbium (III) ions formed in crystalline phases different from Ln2bdc3·4H2O. (TbxLu1−x)2bdc3·nH2O (x = 0–0.01) compounds, synthesized from concentrated solutions, dominantly existed in the Ln2bdc3 crystal structure with CN(Ln3+) = 7. (TbxLu1−x)2bdc3·nH2O (x = 0–0.01) MOFs obtained from diluted solutions formed as Ln2bdc3·10H2O crystalline phases with CN(Ln3+) = 7. At 2–25 at. %, Tb3+ ion binary mixtures of the aforementioned crystalline phases were observed. All of the synthesized samples containing Tb3+ ions demonstrated admirable green luminescence upon 280nm excitation due to the 5D47FJ (J = 3–6) transitions of the Tb3+ ions. Upon UV-photon absorption, terephthalate ion was promoted into the Sn(1ππ*) excited electronic state, which was followed by the fast internal conversion to S1(1ππ*) and then to the T1(3ππ*) triplet electronic excited state via efficient intersystem crossing due to the presence of the heavy lanthanide ion. The T1 state of the terephthalate ion lies slightly higher in energy than the 5D4 level of the Tb3+ ion, resulting in the efficient energy transfer to this level that was followed by radiative 5D47FJ (J = 3–6) transitions. The Tb3+ ions in Ln2bdc3·4H2O, Ln2bdc3·10H2O, and Ln2bdc3·10H2O crystal structures demonstrated different fine structures in their emission bands due to the different local symmetry of the Tb3+ ions in these three types of crystalline structures. The 5D4 excited state lifetimes and photoluminescence quantum yields of (TbxLu1−x)2bdc3 (x = 0.01, 0.5, 0.1) compounds were significantly larger than for samples of (TbxLu1−x)2bdc3·4H2O (x = 0.6, 1) and (TbxLu1−x)2bdc3·10H2O (x = 0.01, 0.5, 0.1) due to the absence of the luminescence quenching of the Tb3+ by coordinated water molecules. Meanwhile, we cannot rule out effect of the crystalline structure on the relative energies of the T1(3ππ*) triplet’s electronic excited state and the 5D4 level of Tb3+ ions, which affect the efficiency of the T1-to-5D4 energy transfer efficiency, resulting in PLQY changes. As a result of our study, we synthesized the material, namely (Tb0.1Lu0.9)2bdc3·1.4H2O, which has one of the highest PLQY among Tb-based MOFs, 95%.

Author Contributions

Conceptualization, A.S.M. and V.G.N.; methodology, A.S.M., Y.N.T. and V.G.N.; validation, M.N.R. and I.E.K.; formal analysis, A.S.M., A.S.P. and V.G.N.; investigation, A.S.M., A.S.P., S.N.O., I.E.K., O.S.B. and V.G.N.; resources, A.S.M., M.Y.S. and N.A.B.; data curation, A.S.M. and V.G.N.; writing—original draft preparation, A.S.M., N.A.B. and V.G.N.; writing—review and editing, M.Y.S., Y.N.T., M.Y.S., I.E.K., N.A.B., V.G.N., S.S.K. and A.S.M.; visualization, A.S.M., V.G.N. 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 22 February 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 conflict of interest.

Sample Availability

Samples of the compounds are available from the authors.

References

  1. Liu, Y.-Q.; Qin, L.-Y.; Li, H.-J.; Wang, Y.-X.; Zhang, R.; Shi, J.-M.; Wu, J.-H.; Dong, G.-X.; Zhou, P. Application of lanthanide-doped upconversion nanoparticles for cancer treatment: A review. Nanomedicine 2021, 16, 2207–2242. [Google Scholar] [CrossRef]
  2. Zhao, L.; Zhang, W.; Wu, Q.; Fu, C.; Ren, X.; Lv, K.; Ma, T.; Chen, X.; Tan, L.; Meng, X. Lanthanide europium MOF nanocomposite as the theranostic nanoplatform for microwave thermo-chemotherapy and fluorescence imaging. J. Nanobiotechnology 2022, 20, 133. [Google Scholar] [CrossRef]
  3. Yin, K.; Wu, S.; Zheng, H.; Gao, L.; Liu, J.; Yang, C.; Qi, L.W.; Peng, J. Lanthanide metal-organic framework-based fluorescent sensor arrays to discriminate and quantify ingredients of natural medicine. Langmuir 2021, 37, 5321–5328. [Google Scholar] [CrossRef]
  4. Lustig, W.P.; Mukherjee, S.; Rudd, N.D.; Desai, A.V.; Li, J.; Ghosh, S.K. Metal–organic frameworks: Functional luminescent and photonic materials for sensing applications. Chem. Soc. Rev. 2017, 46, 3242–3285. [Google Scholar] [CrossRef]
  5. Zhang, Y.; Liu, S.; Zhao, Z.-S.; Wang, Z.; Zhang, R.; Liu, L.; Han, Z.-B. Recent progress in lanthanide metal–organic frameworks and their derivatives in catalytic applications. Inorg. Chem. Front. 2021, 8, 590–619. [Google Scholar] [CrossRef]
  6. Vahedigharehchopogh, N.; Kıbrıslı, O.; Erol, E.; Çelikbilek Ersundu, M.; Ersundu, A.E. A straightforward approach for high-end anti-counterfeiting applications based on NIR laser-driven lanthanide doped luminescent glasses. J. Mater. Chem. C Mater. 2021, 9, 2037–2046. [Google Scholar] [CrossRef]
  7. Kaczmarek, A.M.; Liu, Y.Y.; Wang, C.; Laforce, B.; Vincze, L.; van der Voort, P.; van Hecke, K.; van Deun, R. Lanthanide “chameleon” multistage anti-counterfeit materials. Adv. Funct. Mater. 2017, 27, 1700258. [Google Scholar] [CrossRef]
  8. Liu, D.; Lu, K.; Poon, C.; Lin, W. Metal–organic frameworks as sensory materials and imaging agents. Inorg. Chem. 2014, 53, 1916–1924. [Google Scholar] [CrossRef] [PubMed]
  9. Amoroso, A.J.; Pope, S.J.A. Using lanthanide ions in molecular bioimaging. Chem. Soc. Rev. 2015, 44, 4723–4742. [Google Scholar] [CrossRef] [Green Version]
  10. Zhang, P.; Liang, L.; Liu, X. Lanthanide-doped nanoparticles in photovoltaics—More than just upconversion. J. Mater. Chem. C Mater. 2021, 9, 16110–16131. [Google Scholar] [CrossRef]
  11. Amirkhanov, V.M.; Vishnevsky, D.G.; Ovdenko, V.N.; Chuprina, N.G.; Mokrinskaya, E.V.; Zozulya, V.A.; Shatrava, Y.O.; Ovchinnikov, V.A.; Sliva, T.Y.; Mel’nik, A.K.; et al. Photovoltaic properties of polymer composites doped with binuclear lanthanide complexes derived from 3,6-Dipyridin-2-YL-1,2,4,5-Tetrazine with carbacylamidophosphate ligands. J. Appl. Spectrosc. 2021, 87, 1135–1140. [Google Scholar] [CrossRef]
  12. Chen, D.; Wang, Y.; Hong, M. Lanthanide nanomaterials with photon management characteristics for photovoltaic application. Nano Energy 2012, 1, 73–90. [Google Scholar] [CrossRef]
  13. Ayscue, R.L.; Verwiel, C.P.; Bertke, J.A.; Knope, K.E. Excitation-dependent photoluminescence color tuning in lanthanide-organic hybrid materials. Inorg. Chem. 2020, 59, 7539–7552. [Google Scholar] [CrossRef]
  14. Binnemans, K. Lanthanide-based luminescent hybrid materials. Chem. Rev. 2009, 109, 4283–4374. [Google Scholar] [CrossRef] [Green Version]
  15. Cui, Y.; Yue, Y.; Qian, G.; Chen, B. Luminescent functional metal–organic frameworks. Chem. Rev. 2012, 112, 1126–1162. [Google Scholar] [CrossRef]
  16. 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]
  17. Massi, M.; Ogden, M. Luminescent lanthanoid calixarene complexes and materials. Materials 2017, 10, 1369. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  18. Tu, X.; Tao, Y.; Chen, J.; Du, C.; Jin, Q.; He, Y.; Yang, J.; Huang, S.; Chen, W. Dipicolinic Acid-Tb3+/Eu3+ lanthanide fluorescence sensor array for rapid and visual discrimination of botanical origin of honey. Foods 2022, 11, 3388. [Google Scholar] [CrossRef] [PubMed]
  19. Alpha, B.; Ballardini, R.; Balzani, V.; Lehn, J.-M.; Perathoner, S.; Sabbatini, N. antenna effect in luminescent lanthanide cryptates: A photophysical study. Photochem. Photobiol. 1990, 52, 299–306. [Google Scholar] [CrossRef]
  20. Utochnikova, V.V.; Kuzmina, N.P. Photoluminescence of lanthanide aromatic carboxylates. Russ. J. Coord. Chem. /Koord. Khimiya 2016, 42, 679–694. [Google Scholar] [CrossRef]
  21. Zhou, X.; Wang, H.; Jiang, S.; Xiang, G.; Tang, X.; Luo, X.; Li, L.; Zhou, X. Multifunctional luminescent material Eu(III) and Tb(III) complexes with Pyridine-3,5-Dicarboxylic acid linker: Crystal structures, tunable emission, energy transfer, and temperature sensing. Inorg. Chem. 2019, 58, 3780–3788. [Google Scholar] [CrossRef]
  22. Orlova, A.V.; Kozhevnikova, V.Y.; Lepnev, L.S.; Goloveshkin, A.S.; Le-Deigen, I.M.; Utochnikova, V.V. NIR Emitting Terephthalates (Sm Dy Gd1--)2(Tph)3(H2O)4 for Luminescence Thermometry in the Physiological Range. Journal of Rare Earths 2020, 38, 492–497. [Google Scholar] [CrossRef]
  23. 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] [Green Version]
  24. Zhao, D.; Yue, D.; Zhang, L.; Jiang, K.; Qian, G. Cryogenic luminescent Tb/Eu-MOF thermometer based on a fluorine-modified tetracarboxylate ligand. Inorg. Chem. 2018, 57, 12596–12602. [Google Scholar] [CrossRef] [PubMed]
  25. Feng, T.; Ye, Y.; Liu, X.; Cui, H.; Li, Z.; Zhang, Y.; Liang, B.; Li, H.; Chen, B. A robust mixed-lanthanide PolyMOF membrane for ratiometric temperature sensing. Angew. Chem. 2020, 132, 21936–21941. [Google Scholar] [CrossRef]
  26. Kaur, H.; Sundriyal, S.; Pachauri, V.; Ingebrandt, S.; Kim, K.-H.; Sharma, A.L.; Deep, A. Luminescent metal-organic frameworks and their composites: Potential future materials for organic light emitting displays. Coord. Chem. Rev. 2019, 401, 213077. [Google Scholar] [CrossRef]
  27. Utochnikova, V.V.; Latipov, E.V.; Dalinger, A.I.; Nelyubina, Y.V.; Vashchenko, A.A.; Hoffmann, M.; Kalyakina, A.S.; Vatsadze, S.Z.; Schepers, U.; Bräse, S.; et al. Lanthanide pyrazolecarboxylates for OLEDs and bioimaging. J. Lumin. 2018, 202, 38–46. [Google Scholar] [CrossRef]
  28. Guo, H.; Zhu, Y.; Qiu, S.; Lercher, J.A.; Zhang, H. Coordination modulation induced synthesis of nanoscale Eu1−xTbx-metal-organic frameworks for luminescent thin films. Adv. Mater. 2010, 22, 4190–4192. [Google Scholar] [CrossRef]
  29. Rao, X.; Huang, Q.; Yang, X.; Cui, Y.; Yang, Y.; Wu, C.; Chen, B.; Qian, G. Color tunable and white light emitting Tb3+ and Eu3+ doped lanthanide metal–organic framework materials. J. Mater. Chem. 2012, 22, 3210. [Google Scholar] [CrossRef]
  30. Vialtsev, M.B.; Tcelykh, L.O.; Kozlov, M.I.; Latipov, E.V.; Bobrovsky, A.Y.; Utochnikova, V.V. Terbium and europium aromatic carboxylates in the polystyrene matrix: The first metal-organic-based material for high-temperature thermometry. J. Lumin. 2021, 239, 118400. [Google Scholar] [CrossRef]
  31. Kim, J.H.; Lepnev, L.S.; Utochnikova, V.V. Dual Vis-NIR emissive bimetallic naphthoates of Eu-Yb-Gd: A new approach toward Yb luminescence intensity increase through Eu → Yb energy transfer. Phys. Chem. Chem. Phys. 2021, 23, 7213–7219. [Google Scholar] [CrossRef]
  32. 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]
  33. Rieter, W.J.; Taylor, K.M.L.; An, H.; Lin, W.; Lin, W. Nanoscale metal-organic frameworks as potential multimodal contrast enhancing agents. J. Am. Chem. Soc. 2006, 128, 9024–9025. [Google Scholar] [CrossRef] [Green Version]
  34. Dhananjaya, N.; Nagabhushana, H.; Nagabhushana, B.M.; Rudraswamy, B.; Shivakumara, C.; Narahari, K.; Chakradhar, R.P.S. Enhanced photoluminescence of Gd 2O 3:Eu 3+ nanophosphors with Alkali (M = Li +, Na +, K +) metal ion Co-Doping. Spectrochim. Acta A Mol. Biomol. Spectrosc. 2012, 86, 8–14. [Google Scholar] [CrossRef]
  35. Kumar, D.; Sharma, M.; Pandey, O.P. Effect of Co-Doping metal ions (Li+, Na+ and K +) on the structural and photoluminescent properties of nano-sized Y2O3:Eu3+ synthesized by co-precipitation method. Opt. Mater. 2014, 36, 1131–1138. [Google Scholar] [CrossRef]
  36. Kumari, P.; Manam, J. Enhanced Red Emission on Co-Doping of Divalent Ions (M2+ = Ca2+, Sr2+, Ba2+) in YVO4:Eu3+ Phosphor and spectroscopic analysis for its application in display devices. Spectrochim. Acta A Mol. Biomol. Spectrosc. 2016, 152, 109–118. [Google Scholar] [CrossRef] [PubMed]
  37. Mikalauskaite, I.; Pleckaityte, G.; Skapas, M.; Zarkov, A.; Katelnikovas, A.; Beganskiene, A. Emission spectra tuning of upconverting NaGdF4:20% Yb, 2% Er nanoparticles by Cr3+ Co-doping for optical temperature sensing. J. Lumin. 2019, 213, 210–217. [Google Scholar] [CrossRef]
  38. Du, K.; Xu, X.; Yao, S.; Lei, P.; Dong, L.; Zhang, M.; Feng, J.; Zhang, H. Enhanced upconversion luminescence and controllable phase/shape of NaYF4:Yb/Er crystals through Cu2+ ion doping. Cryst. Eng. Comm. 2018, 20, 1945–1953. [Google Scholar] [CrossRef]
  39. 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]
  40. 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]
  41. 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] [Green Version]
  42. 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] [PubMed]
  43. Yin, P.X.; Zhang, J.; Qin, Y.Y.; Cheng, J.K.; Li, Z.J.; Yao, Y.G. Role of molar-ratio, temperature and solvent on the Zn/Cd 1,2,4-triazolate system with novel topological architectures. Cryst. Eng. Comm. 2011, 13, 3536–3544. [Google Scholar] [CrossRef]
  44. Seetharaj, R.; Vandana, P.V.; Arya, P.; Mathew, S. Dependence of solvents, PH, molar ratio and temperature in tuning metal organic framework architecture. Arab. J. Chem. 2019, 12, 295–315. [Google Scholar] [CrossRef] [Green Version]
  45. Dighe, A.V.; Nemade, R.Y.; Singh, M.R. Modeling and simulation of crystallization of metal–organic frameworks. Processes 2019, 7, 527. [Google Scholar] [CrossRef] [Green Version]
  46. Kofod, N.; Sørensen, T.J. Tb 3+ Photophysics: Mapping excited state dynamics of [Tb(H 2 O) 9 ] 3+ using molecular photophysics. J. Phys. Chem. Lett. 2022, 13, 11968–11973. [Google Scholar] [CrossRef]
  47. Cotton, S.A. Establishing coordination numbers for the lanthanides in simple complexes. Comptes Rendus Chimie 2005, 8, 129–145. [Google Scholar] [CrossRef]
  48. Allen, P.G.; Bucher, J.J.; Shuh, D.K.; Edelstein, N.M.; Craig, I. Coordination chemistry of trivalent lanthanide and actinide ions in dilute and concentrated chloride solutions. Inorg. Chem. 2000, 39, 595–601. [Google Scholar] [CrossRef]
  49. Yeung, H.H.-M.; Sapnik, A.F.; Massingberd-Mundy, F.; Gaultois, M.W.; Wu, Y.; Fraser, D.A.X.; Henke, S.; Pallach, R.; Heidenreich, N.; Magdysyuk, O.V.; et al. Control of metal–organic framework crystallization by metastable intermediate pre-equilibrium species. Angewandte Chemie 2018, 58, 566–571. [Google Scholar] [CrossRef]
  50. Shannon, R.D. Revised effective ionic Radii and systematic studies of interatomic distances in halides and chalcogenides. Acta Crystallographica Section A 1976, 32, 751–767. [Google Scholar] [CrossRef]
  51. Haquin, V.; Etienne, M.; Daiguebonne, C.; Freslon, S.; Calvez, G.; Bernot, K.; le Pollès, L.; Ashbrook, S.E.; Mitchell, M.R.; Bünzli, J.C.; et al. Color and brightness tuning in heteronuclear lanthanide terephthalate coordination polymers. Eur. J. Inorg. Chem. 2013, 2013, 3464–3476. [Google Scholar] [CrossRef] [Green Version]
  52. Huang, C.-H.; Liu, W.-R.; Kuo, T.-W.; Chen, T.-M. A study on the luminescence and energy transfer of green-emitting Ca9Y(PO4)7:Ce3+,Tb3+ phosphor for fluorescent lamp application. Chemistry 2011, 1, 9–15. [Google Scholar] [CrossRef]
  53. Kalusniak, S.; Castellano-Hernández, E.; Yalçinoğlu, H.; Tanaka, H.; Kränkel, C. Spectroscopic properties of Tb3+ as an ion for visible lasers. Appl. Phys. B 2022, 128, 33. [Google Scholar] [CrossRef]
  54. Ofelt, G.S. Intensities of crystal spectra of rare-earth ions. J. Chem. Phys. 1962, 37, 511–520. [Google Scholar] [CrossRef]
  55. Judd, B.R. Optical absorption intensities of rare-earth ions. Phys. Rev. 1962, 127, 750–761. [Google Scholar] [CrossRef]
  56. 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]
  57. 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]
  58. Hölsä, J.; Leskelä, M.; Niinistö, L. Concentration quenching of Tb3+ luminescence in LaOBr and Gd2O2S phosphors. Mater. Res. Bull. 1979, 14, 1403–1409. [Google Scholar] [CrossRef]
  59. Zhang, W.; Kou, H.; Ge, L.; Zhang, Y.; Lin, L.; Li, W. Effects of Doping Ions on the Luminescence Performance of Terbium Doped Gadolinium Polysulfide Phosphor. J Phys Conf Ser 2020, 1549, 032064. [Google Scholar] [CrossRef]
  60. Kolesnikov, I.E.; Kalinichev, A.A.; Kurochkin, M.A.; Golyeva, E.V.; Terentyeva, A.S.; Kolesnikov, E.Y.; Lähderanta, E. Structural, luminescence and thermometric properties of nanocrystalline YVO 4:Dy3+ temperature and concentration series. Sci. Rep. 2019, 9, 1–14. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  61. Kolesnikov, I.E.; Mamonova, D.V.; Lähderanta, E.; Kurochkin, A.V.; Mikhailov, M.D. The impact of doping concentration on structure and photoluminescence of Lu2O3:Eu3+ nanocrystals. J. Lumin. 2017, 187, 26–32. [Google Scholar] [CrossRef]
  62. Mao, Z.Y.; Zhu, Y.C.; Zeng, Y.; Gan, L.; Wang, Y. Concentration quenching and resultant photoluminescence adjustment for Ca3Si2O7:Tb3+ green-emitting phosphor. J. Lumin. 2013, 143, 587–591. [Google Scholar] [CrossRef]
  63. Ivanova, A.A.; Gontcharenko, V.E.; Lunev, A.M.; Sidoruk, A.V.; Arkhipov, I.A.; Taydakov, I.V.; Belousov, Y.A. New carboxylate anionic Sm-MOF: Synthesis, structure and effect of the isomorphic substitution of Sm3+ with Gd3+ and Tb3+ Ions on the luminescent properties. Inorganics 2022, 10, 104. [Google Scholar] [CrossRef]
  64. Schwarzenbach, G.; Flashka, H. Complexometric Titrations, 2nd ed.; Methuen: London, UK, 1969. [Google Scholar]
Figure 1. The PXRD patterns of (TbxLu1−x)2bdc3·nH2O (x = 0–1) MOFs synthesized from the diluted (a) and concentrated (b) solutions as well as the PXRD patterns of Tb2bdc3·4H2O [40], Lu2bdc3·10H2O [41], and Tb2bdc3 [42] simulated from the single-crystal structures.
Figure 1. The PXRD patterns of (TbxLu1−x)2bdc3·nH2O (x = 0–1) MOFs synthesized from the diluted (a) and concentrated (b) solutions as well as the PXRD patterns of Tb2bdc3·4H2O [40], Lu2bdc3·10H2O [41], and Tb2bdc3 [42] simulated from the single-crystal structures.
Molecules 28 02378 g001
Figure 2. The crystal structures of Tb2bdc3·4H2O (a), Tb2bdc3 (b), and Lu2bdc3·10H2O (c) generated from the single-crystal diffraction data [40,41,42].
Figure 2. The crystal structures of Tb2bdc3·4H2O (a), Tb2bdc3 (b), and Lu2bdc3·10H2O (c) generated from the single-crystal diffraction data [40,41,42].
Molecules 28 02378 g002
Figure 3. Thermogravimetric analysis (TGA) curves showing the weight loss profile of (TbxLu1−x)2bdc3·nH2O materials synthesized from the diluted (a) and concentrated (b) solutions during thermal decomposition.
Figure 3. Thermogravimetric analysis (TGA) curves showing the weight loss profile of (TbxLu1−x)2bdc3·nH2O materials synthesized from the diluted (a) and concentrated (b) solutions during thermal decomposition.
Molecules 28 02378 g003
Figure 4. The number of water molecules per one formula unit calculated from TGA data for (TbxLu1−x)2bdc3·nH2O materials synthesized from the diluted (a) and concentrated (b) solutions; the molar fractions of Ln2bdc3·4H2O and Ln2bdc3·10H2O (c) and the molar fractions of Ln2bdc3·4H2O and Ln2bdc3 (d) as functions of Tb3+ concentration for (TbxLu1−x)2bdc3·nH2O materials synthesized from the diluted and concentrated solutions, respectively.
Figure 4. The number of water molecules per one formula unit calculated from TGA data for (TbxLu1−x)2bdc3·nH2O materials synthesized from the diluted (a) and concentrated (b) solutions; the molar fractions of Ln2bdc3·4H2O and Ln2bdc3·10H2O (c) and the molar fractions of Ln2bdc3·4H2O and Ln2bdc3 (d) as functions of Tb3+ concentration for (TbxLu1−x)2bdc3·nH2O materials synthesized from the diluted and concentrated solutions, respectively.
Molecules 28 02378 g004
Figure 5. Emission spectra of (TbxLu1−x)2bdc3·nH2O materials synthesized from the diluted (a) and concentrated (b) solutions upon 280-nm excitation at the selected Tb3+ concentrations.
Figure 5. Emission spectra of (TbxLu1−x)2bdc3·nH2O materials synthesized from the diluted (a) and concentrated (b) solutions upon 280-nm excitation at the selected Tb3+ concentrations.
Molecules 28 02378 g005
Figure 6. The photoluminescence decay curves of (TbxLu1−x)2bdc3·nH2O materials synthesized from the diluted (a) and concentrated (b) solutions upon UV-excitation at the selected Tb3+ concentrations.
Figure 6. The photoluminescence decay curves of (TbxLu1−x)2bdc3·nH2O materials synthesized from the diluted (a) and concentrated (b) solutions upon UV-excitation at the selected Tb3+ concentrations.
Molecules 28 02378 g006
Table 1. Lifetimes (τ) and photoluminescence quantum yields (ΦPL) of (TbxLu1−x)2bdc3·nH2O materials at the selected Tb3+ concentrations synthesized from the diluted (Series 1) and concentrated (Series 2) solutions.
Table 1. Lifetimes (τ) and photoluminescence quantum yields (ΦPL) of (TbxLu1−x)2bdc3·nH2O materials at the selected Tb3+ concentrations synthesized from the diluted (Series 1) and concentrated (Series 2) solutions.
Series 1 (From Diluted Solutions)Series 2 (From Concentrated Solutions)
ΧTb (at. %)τ, msΦPL, %ΧTb (at. %)τ1, msτ2, msΦPL, %
11.12 ± 0.023811.17 ± 0.042.63 ± 0.1077
51.12 ± 0.025651.53 ± 0.053.00 ± 0.2488
101.08 ± 0.0158101.02 ± 0.032.61 ± 0.0895
600.92 ± 0.0260600.94 ± 0.02 60
1000.70 ± 0.01491000.69 ± 0.01 49
Table 2. The volumes of the initial TbCl3 and LuCl3 solutions used for the synthesis of (TbxLu1−x)2bdc3·nH2O MOFs.
Table 2. The volumes of the initial TbCl3 and LuCl3 solutions used for the synthesis of (TbxLu1−x)2bdc3·nH2O MOFs.
ΧTb (at. %)V(0.2M TbCl3), mLV(0.2M LuCl3), mL
00.002.00
10.021.98
50.101.90
100.201.80
150.301.70
200.401.60
250.501.50
300.601.40
601.200.80
1002.000.00
Table 3. Tb3+ atomic fractions (relative to the total amount of Tb3+ and Lu3+) in (TbxLu1−x)2bdc3·nH2O compounds synthesized from the diluted (Series 1) and concentrated (Series 2) solutions. Measurements were taken during synthesis and obtained from EDX data.
Table 3. Tb3+ atomic fractions (relative to the total amount of Tb3+ and Lu3+) in (TbxLu1−x)2bdc3·nH2O compounds synthesized from the diluted (Series 1) and concentrated (Series 2) solutions. Measurements were taken during synthesis and obtained from EDX data.
Series 1 (from Diluted Solutions)Series 2 (from Concentrated Solutions)
Χtb (At. %), TakenΧTb (%), EDXΧTb (at. %), TakenΧTb (%), EDX
0000
10.74 ± 0.0710.70 ± 0.07
54.6 ± 0.554.6 ± 0.5
109 ± 11010 ± 1
1515 ± 31514 ± 1
2019 ± 22020 ± 2
2526 ± 32523 ± 2
3029 ± 33027 ± 3
6057 ± 56057 ± 5
100100100100
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

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. https://doi.org/10.3390/molecules28052378

AMA Style

Nosov VG, Toikka YN, Petrova AS, Butorlin OS, Kolesnikov IE, Orlov SN, Ryazantsev MN, Kolesnik SS, Bogachev NA, Skripkin MY, et al. Brightly Luminescent (TbxLu1−x)2bdc3·nH2O MOFs: Effect of Synthesis Conditions on Structure and Luminescent Properties. Molecules. 2023; 28(5):2378. https://doi.org/10.3390/molecules28052378

Chicago/Turabian Style

Nosov, Viktor G., Yulia N. Toikka, Anna S. Petrova, Oleg S. Butorlin, Ilya E. Kolesnikov, Sergey N. Orlov, Mikhail N. Ryazantsev, Stefaniia S. Kolesnik, Nikita A. Bogachev, Mikhail Yu. Skripkin, and et al. 2023. "Brightly Luminescent (TbxLu1−x)2bdc3·nH2O MOFs: Effect of Synthesis Conditions on Structure and Luminescent Properties" Molecules 28, no. 5: 2378. https://doi.org/10.3390/molecules28052378

Article Metrics

Back to TopTop