Next Article in Journal
Thin-Film Composite Polyamide Membranes Modified with HKUST-1 for Water Treatment: Characterization and Nanofiltration Performance
Previous Article in Journal
Effects of the Size and Loading of Chrome-Tanned Leather Shavings on the Properties of Styrene–Butadiene Rubber Compounds
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Engineering Photoluminescence of Lanthanide Doped Yttrium-MOF-76 for Volatile Organic Compound Sensing

1
Área de Química General e Inorgánica “Dr. G. F. Puelles”, Facultad de Química, Bioquímica y Farmacia, Universidad Nacional de San Luis, Ejército de los Andes 950, San Luis 5700, Argentina
2
Instituto de Investigaciones en Tecnología Química (INTEQUI), Almirante Brown 1455, San Luis 5700, Argentina
3
Área Química, Instituto de Ciencias, Universidad Nacional de General Sarmiento, CONICET. J. M. Gutiérrez 1150, Buenos Aires 1613, Argentina
4
Instituto Multidisciplinario de Investigaciones Biológicas de San Luis (IMIBIO-SL), CONICET-UNSL, Av. Ejército de los Andes 950, San Luis 5700, Argentina
5
L3—Luminescent Lanthanide Lab, Department of Chemistry, Ghent University, Krijgslaan 281, Building S3, 9000 Gent, Belgium
*
Author to whom correspondence should be addressed.
Polymers 2025, 17(9), 1135; https://doi.org/10.3390/polym17091135
Submission received: 29 January 2025 / Revised: 12 April 2025 / Accepted: 18 April 2025 / Published: 22 April 2025

Abstract

:
A set of three-dimensional metal-organic frameworks, named MOF-76, belonging to the tetragonal P4322 space group, based on [Y(BTC)(H2O)](DMF)1.1 (1,3,5-benzenetricarboxylate) doped with Eu3+, Tb3+, and Eu3+/Tb3+ were obtained under solvothermal conditions and fully characterized by powder X-ray diffraction, thermal, and vibrational analyses. In addition, upon UV light excitation (280 nm), all the powdered samples exhibited fine 4f-4f transitions, of which the 5D07F2 (Eu3+) and 5D47F5 (Tb3+) were the most intense ones. All samples were photophysically analyzed by determining the luminescence lifetimes, and their emission colors were quantified by calculating their chromaticities and color purities. Moreover, the intrinsic quantum yield, radiative, and non-radiative constants were calculated and compared to establish a structure–property relationship. Specifically, the Eu/Tb co-doped sample was employed to monitor its hypersensitive emissions in the presence of small volatile organic compounds (VOCs), showing quenching or enhancement of emission in protic and non-protic solvents. Furthermore, DFT calculations were carried out to understand the energy transfer processes between the sensor and the respective analytes. These results are promising for the development of solid-state lighting devices and colorimetric chemical sensors for specific compounds.

1. Introduction

Metal–organic frameworks (MOFs) are a type of coordination polymer that have attracted significant attention due to their diverse applications, including ion exchange [1,2], gas separation [3], heterogeneous catalysis [4], optoelectronics [5], optomagnetism [6], controlled drug release [7], and antimicrobial composites for human health [8]. These materials stand out for their high porosity, large surface area, and structural flexibility, enabling the design of micro- and nanomaterials with tailored properties derived from metal ions, organic linkers, and their synergistic interactions [9,10,11].
Among MOFs, lanthanide-based metal–organic frameworks (Ln-MOFs) are particularly valued for photonics and magnetism, which are attributed to the unique 4f-4f electronic configuration. The 4f-4f transitions result in sharp spectral lines, varied lifetimes, and high quantum yields, making Ln-MOFs ideal for applications such as phosphors, lasers, optical amplifiers, solid-state lighting, full-color displays, and backlighting systems [5,12,13,14]. Moreover, there are numerous examples of Ln-MOFs employed as models for water remediation by photocatalysis [15].
Furthermore, mixed-lanthanide MOFs offer enhanced functionality compared to single-lanthanide counterparts, enabling tunable white light emission and efficient temperature sensing, thereby broadening their potential applications in fields requiring precise optical responses [3,7,16].
Within Ln-MOFs, MOF-76 stands out for its remarkable structural and optical properties. This structure, based on 1,3,5-benzenetricarboxylate (BTC3−) ligands coordinated to lanthanide centers, crystallizes in the tetragonal P4322 (#95) space group and presents well-defined porous 1D channel structure cavities [17,18,19]. Additionally, Eu-MOF-76 and Tb-MOF-76 exhibit notable 4f-4f emissions, such as red emission at 610 nm (Eu3⁺) and green emission at 540 nm (Tb3⁺), which are further amplified by the “antenna effect”, where organic ligands efficiently transfer energy to lanthanide ions, enhancing their photoluminescence intensity [18]. These attributes make MOF-76 a promising candidate for applications in chemical and thermal sensors, photocatalysis, and light-emitting devices [20].
One of the most pressing applications of MOF materials lies in the detection and recognition of volatile organic compounds (VOCs), which are critical for environmental monitoring, industrial safety, and public health [21]. VOCs are emitted from a plethora of industrial processes, household products, and natural sources, contributing to air pollution and posing serious health risks, including respiratory issues, neurological disorders, and even cancer with prolonged exposure [22]. Additionally, VOCs play a crucial role in atmospheric chemistry, participating in reactions that lead to the formation of ground-level ozone and secondary organic aerosols, which exacerbate air quality problems [23].
Despite the urgent need for effective VOC sensing technologies, existing detection methods face significant challenges. Traditional techniques such as gas chromatography and mass spectrometry provide high sensitivity and selectivity; however, they are often expensive, time-consuming, and require specialized instrumentation, limiting their practicality for real-time or on-site monitoring [24,25,26,27]. Moreover, conventional solid-state sensors, including metal oxide and electrochemical sensors, frequently suffer from poor selectivity, slow response and recovery times, and sensitivity to environmental factors such as humidity and temperature fluctuations, leading to inconsistent performance [28]. These limitations highlight the need for alternative sensing platforms that combine high sensitivity, selectivity, rapid response, and recyclability.
MOFs have demonstrated exceptional potential as VOC sensors due to their high porosity, selective adsorption capabilities, chemical and thermal stabilities, and tunable luminescent properties [29,30]. The ability of luminescent MOFs to interact with VOCs through host–guest interactions enables highly sensitive and selective detection, as analyte adsorption can modulate photoluminescence properties through enhancement or quenching, wavelength shifts, or lifetime variations, providing a sensitive and selective detection mechanism [31]. However, achieving precise control over these luminescent responses and understanding the fundamental mechanisms governing them remain ongoing challenges.
In this study, we explore a novel application of MOF-76 doped with europium and terbium for solvent-specific sensing devices. Powder X-ray diffraction, thermogravimetric analysis, differential scanning calorimetry, and infrared spectroscopy were employed to assess the material’s structural, vibrational, and thermal properties. An in-depth photophysical characterization was carried out by recording excitation and emission spectra, calculating the lifetime values, and europium’s intrinsic quantum yields. Also, the lanthanide energy transfer and color quantification were estimated. Finally, the luminescent response of the Eu/Tb co-doped MOF-76 in various solvents was evaluated, demonstrating its potential for developing advanced chemical sensors capable of detecting VOCs.

2. Materials and Methods

2.1. Synthesis

All reagents and solvents were used as received from Sigma-Aldrich (Burlington, MA, USA) without further purification: EuCl3∙6H2O, TbCl3∙6H2O, YCl3.6H2O, N,N′-dimethylformamide (DMF), and 1,3,5-benzenetricarboxylic acid (H3BTC, H6C9O6). The crystalline materials were obtained as crystalline solids under solvothermal conditions using 43 Parr reactors. Compounds with formula [Ln(C9H3O4)(H2O)]∙(DMF)1.1 (further labeled as Ln-BTC) were obtained following previously reported procedures with slight modifications [32,33].

2.1.1. Y-BTC

H3BTC (0.60 mmol, 0.1208 g) and YCl3·6H2O (0.50 mmol, 0.1517 g) were dissolved in DMF (9 mL) and H2O (3 mL) at room temperature. The mixture was stirred for 30 min, then transferred to a Teflon-lined Parr reactor and heated at 80 °C for 72 h. After that, the product was filtered and washed with 10 mL of water and 10 mL of DMF. Finally, the crystalline product was dried at room temperature for 48 h.

2.1.2. Ln@Y-BTC

For the synthesis of lanthanide-doped compounds, a similar procedure to Y-BTC was followed, except for the addition of 5% of EuCl3·6H2O (0.0201 g) (Eu@Y-BTC) or TbCl3·6H2O (0.0204 g) (Tb@Y-BTC). For co-doped samples, Eu2.5Tb2.5@Y-BTC was prepared using the same procedure explained before, but including EuCl3·6H2O in a 2.5% (0.0101 g) and 2.5% TbCl3·6H2O (0.0104 g) amount. Also, Eu1.25Tb3.75@Y-BTC was prepared by incorporating 1.25% EuCl3·6H2O (0.050 g) and 2.5% TbCl3·6H2O (0.0154 g). The incorporation of Eu and Tb into the mixed MOFs was confirmed by ICP analysis. As shown in Table S4, the measured percentages of the dopant elements are slightly lower than the nominal amounts used in the synthesis. However, the Eu:Tb ratio remains consistent with the intended doping ratio, indicating a uniform incorporation of both lanthanides into the frameworks.

2.2. Characterization

The powder X-ray diffraction (PXRD) plots were recorded with a Rigaku–Ultima IV type II diffractometer. A scanning step of 0.05° into the 5–50 2-theta Bragg angles range with an exposure time of 5 s per step was used to obtain the best counting statistics. Fourier transform infrared (FTIR) spectra were recorded with a Nicolet Protégé 460 spectrometer in the 4000–400 cm−1 range with 64 scans and a spectral resolution of 4 cm−1 by the KBr pellet technique. Thermogravimetric analysis (TGA) was performed using a Shimadzu TGA-51 (Shimadzu Corp., Kyoto, Japan) apparatus under flowing air at a flow rate of 50 mL∙min−1 and a heating rate of 10 °C.min−1. Differential thermal analysis (DTA) was performed with a DSC-50 under air flow at a rate of 50 mL∙min−1 and a heating rate of 10 °C∙min−1.

2.3. Photophysical Characterization and Sensing Assays

Solid-State Luminescence Measurements: The steady-state and time-resolved luminescence measurements were performed using an Edinburgh Instruments FLSP920 spectrometer (Edinburgh Instruments Ltd, Livingston, UK) setup, using a 450 W xenon lamp as the steady-state excitation source and a 60 W pulsed xenon lamp as the time-resolved excitation source (operating at a pulse frequency of 100 Hz). The emission was detected by a Hamamatsu R928P PMT photomultiplier tube (Hamamatsu Co., Shizuoka, Japan). Excitation spectra were corrected for the xenon lamp emission profile, whereas emission spectra were corrected for the detector response curve. All measurements were carried out at a step size of 0.1 nm. Commission Internationale de l’Eclairage (CIE) (x,y) color coordinates were calculated using the Matlab (Version R2020b) program.

2.3.1. Chemical Sensor Studies

The sensing activity of Eu1.25Tb3.75@Y-BTC was investigated by monitoring the emission spectra at 613 nm when exciting the samples at 280 nm. A quartz cuvette with a 1 cm optical path length was employed. The VOC@Eu1.25Tb3.75@Y-BTC suspensions were prepared by introducing 0.2 mg of powdered sample into 4 mL (0.05 mg∙mL−1) of each solvent [bidistilled water, methanol (Carl Roth, Karlsruhe, Germany, ≥99%), acetonitrile (Sigma Aldrich, Burlington, MA, USA, ≥99.9%) ethanol (Fischer Chemical, Pittsburgh, PA, USA, 99.9%), dimethylformamide (Sigma Aldrich, 99.8%), acetone (Acros Organics, Pittsburgh, PA, USA, pure), chloroform (Sigma Aldrich, ≥99.8%), 1,3,5-trimethylbenzene (1,3,5-TMB) (Sigma Aldrich, 98%), and toluene (Sigma Aldrich, 99.9%). The samples were previously ultrasonicated for 30 min.

2.3.2. DFT Calculations

Quantum mechanical calculations were carried out using density-functional theory (DFT) and time-dependent DFT (TDDFT) with the Gaussian 16 software package [34]. Equilibrium geometries of the triplet states were optimized under tight convergence criteria using the range-separated hybrid wB97XD functional [35]. (ESI Theoretical calculations section, Listing S1–S9). This functional, which incorporates long-range corrections and empirical dispersion terms, was selected for its ability to accurately model both electronic structure and non-covalent interactions [36]. The split-valence, triple-zeta basis set augmented with diffuse and polarization functions, 6-311++G(d,p), was used to balance computational efficiency with accuracy [37]. Vibrational frequency analyses confirmed the absence of imaginary frequencies in the optimized geometries, verifying that these structures represent true local minima and ensuring the reliability of the computed results [38] (SI Theoretical calculations section, Table S1). For excited-state calculations, the Tamm–Dancoff approximation (TDA) was employed to provide a robust framework for modelling electronic excitations, particularly in systems where single excitations dominate [39]. Analyte effects were accounted for using the polarizable continuum model (PCM) in TDDFT calculations [40]. A full population analysis was conducted to quantify charge distributions using frontier molecular orbital (FMO) theory (ESI Theoretical calculations section, Figures S1–S9). The excitation assignments were obtained from the TDDFT-derived transition densities analysis [41].

3. Results and Discussion

3.1. Synthesis

The described synthetic procedures led to crystalline products (Figure 1), which were fully characterized by PXRD, TGA-DSC, and FTIR techniques. According to optical microscope observations on the Y-BTC sample, long prismatic block crystals were obtained (Figure 1a). Additionally, a brief structural description of Ln-BTC compounds is presented. The Ln-BTC structure is 3D, belonging to the tetragonal P4322 space group. The asymmetric unit is composed of one hepta-coordinated trivalent lanthanide ion, one BTC3− linker, and one water molecule. Each lanthanide ion is surrounded by six oxygen atoms belonging to carboxylate groups and one oxygen atom from a water molecule (Figure 1b). Moreover, the metallic chain polyhedra are developed in a helical fashion along the c axis (Figure 1c). The chains are linked by BTC3− ligands along the a and b axes, giving rise to a 3D framework. The structure contains unidimensional channels along the c direction, with a circular area of 36 Å2. The incorporation of Eu and Tb into the MOFs was confirmed by ICP-AES analysis. As shown in Table S1, the measured percentage of the dopant elements is slightly lower than the nominal amounts used in the synthesis. However, the Eu:Tb ratio remains consistent with the intended doping ratio, indicating a uniform incorporation of both lanthanides into the framework. Finally, from the topological point of view, the MOF-76 structure can be simplified into bars and dots to get the underlying net. According to Rosi et al. [17], MOF-76 belongs to the pcu-type net classification. This geometry consists of rods packed in a tetragonal fashion, resulting in square channels in the c direction, which are filled with DMF molecules. However, the rods themselves are on 41 helices, but, because of the tritopic nature of the organic SBU, this results in a rather complicated overall topology.
The characterization of Ln@Y-BTC samples by the PXRD technique revealed the isostructural nature of the reported family of compounds in comparison to the simulated pattern from the .cif file of the pristine Y-BTC structure (Figure 2a) [33]. The incorporation of Eu3+ and/or Tb3+ did not represent an alteration of the crystalline structure, as can be seen in the corresponding powder patterns of the sets.
Thermal analysis of the Eu@Y-BTC sample revealed notable stability up to 500 °C, at which point the formation of Ln₂O₃ takes place (see Figure S10). The first mass loss of 5.3% (calculated: 4.56%) in the TGA diagram corresponds to the removal of one coordinated water, accompanied by an endothermic peak in the DSC plot at 105 °C. The second and third mass loss steps, involving the removal of 1.1 DMF molecules with a mass loss of 19.7% (calculated: 20.4%), were accompanied by endothermic signals at 176 °C and 320 °C in the DSC diagram. Finally, the decomposition of the organic moieties is evidenced by a significant exothermic signal in the DSC curve at around 395 °C, and in the range of 490–550 °C, the lanthanide oxide is formed (Figure 2b). Based on these results, the final stoichiometry of Eu@Y-BTC was determined to be [Y0.95Eu0.05(BTC)(H2O)](DMF)1.1. Through vibrational characterization (Figure S11), it was possible to identify bands related to the asymmetric and symmetric modes of the carboxylate groups (1615, 1440, and 1380 cm−1) from the ligand, coordinated water (3400 cm−1), and guest DMF molecules (3068, 2926, and 2860 cm−1).

3.2. Photophysical Studies

Photoluminescence (PL) characterization includes key parameters [42] such as (a) PL spectra; (b) quantum yields; and/or (c) observed luminescence lifetimes (τobs). Moreover, the precise quantification of the emitting light is essential for the development of optical devices used in electronics, and chemical and physical sensors [43]. In this context, the room-temperature solid-state photoluminescence properties of Y-BTC and Ln@Y-BTC compounds were explored. When the Y-BTC is excited at 280 nm, an emission is achieved that corresponds to a broad band located at 430 nm (Figure 3b). This transition is dominant and responsible for the purplish-blue emission, as confirmed by the calculated CIE(x,y) chromaticities coordinates (see Figure 4a). Additionally, the τobs of emission is 0.0216 ms, the smallest value from the sets, which is consistent with typical values for organic molecules.
The photoluminescence behavior of the doped samples, as expected from the lanthanide 4f-transitions, was characterized by recording excitation profiles by monitoring the 4f emitting lines: 613 nm for Eu3⁺ and 544 nm for Tb3⁺. These profiles revealed an intense absorption transition centered at 280 nm (Figure 3a). This transition is seen as a typical broadband, is related to π*← π or π*← n transitions from the organic BTC3− moieties and was selected for setting the excitation wavelength for the sample sets. Upon ligand sensitization (τexc = 280 nm) Eu@Y-BTC exhibits a dual emission from the BTC3− ligand at 439 nm due to π* → n/π* → π transitions, accompanied by fine 4f-transitions centered at 5D07Fn (n = 1–4) transitions observed at 589, 616, 652, and 700 nm, respectively (Figure 3c). The combination of the mentioned transitions led to a purplish-pink emission (Figure 4a). The 5D07F2 transition was the most intense with a τobs of 0.28 ms. In most cases, the hypersensitive 5D07F2 transition is the most intense transition responsible for the typical red emission in Eu-containing compounds [33].
Similarly, Tb@Y-BTC, upon excitation at 280 nm, displayed the characteristic, 5D47Fn (n = 6–0) transitions centered at 487, 544, 580, 622, 649, 673 and 682, which are attributed to Tb3+ 4f*-4f transitions, respectively. In this case, the strongest emission is associated with the 5D47F5 transition (544 nm), which is responsible for the bright green emission (see Figure 3d) [44]. Also, the lifetime value was calculated as 1.36 ms, being the longest among the studied samples in this work. For the Eu2.5Tb2.5@Y-BTC sample, upon 280 nm excitation, two sets of lanthanide transitions belonging to Eu3+ and Tb3+ were observed, yielding an orange-pink emission (Figure 3e). The dominant transition is the 5D07F2 from Eu3+ ions, principally due to the efficient energy migration from the emitting level of Tb3+ to Eu3+. In this sense, the close proximity of the excited levels for both lanthanides makes a metal-to-metal charge transfer feasible [45]. The τobs values for the Eu2.5Tb2.5@Y-BTC sample were calculated from decay data, being 0.24 ms for Eu3+ and 0.28 ms for Tb3+. These lifetime values reinforce the mechanism of energy transfer among the lanthanide ions.
In the Eu1.25Tb3.75@Y-BTC sample, a higher Tb3⁺ content resulted in an increase in the intensity of the 5D47F5 (Tb3⁺) transition (Figure 3f). However, the emission from europium ions remained more prominent, leading to a yellow color. The corresponding τobs values were 0.39 for Eu3+ ions and 0.64 ms for Tb3+, also reinforcing the mechanism of metal-to-metal charge transfer among both metal ions. Additionally, according to the proposed Jabloski diagram (Figure 4b), the emissive level of terbium is higher than that of europium, therefore, it is reasonable to assume a depopulation from the 5D4 (Tb3+) levels to partially feed the 5D0 (Eu3+) levels. Previously, we explored the optical response of RGB-SURMOFs (Red-Green-Blue) devices based on the Gd-BTC structure combined with Eu- and Tb-BTC nanolayers with a heteroepitaxial geometry [35] In that study, an efficient energy transfer was confirmed involving BTC3− → Tb3+ → Eu3+ energy migration, resulting in white light emission. To gain deeper insights into the energy transfer processes in the mixed samples, the energy transfer efficiencies (ηT) from Tb3+ to Eu3+ were calculated using the following formula (Equation (1)) [46]:
η T b E u = 1 τ τ 0
where η T b E u is the energy transfer efficiency and τ0 and τ are the observed lifetimes of Tb3+ ions in the absence and presence of Eu3+ ions, respectively. Thus, the relationship between the energy-transfer efficiency and activator concentration of Eu3+ ions can be analyzed. The value of η T b E u reaches a maximum of 82% in the Eu2.5Tb2.5@Y-BTC sample and a value of 53% in Eu1.25Tb3.75@Y-BTC.
High-quality light performance requires the calculation of the Commission International de l’Eclairage (CIE) x,y chromaticity coordinates and the correlated color temperature (CCT), which are critical parameters for solid-state lighting applications. Quantifying the color emission of different luminescent sources allows their comparison by studying the corresponding light-emitting performance. In this context, the color coordinates are usually calculated using the CIE x,y chromaticity system and plotted in a two-dimensional diagram, providing a visual representation of emission features. The color emission of Y-BTC and Ln@Y-BTC was quantified as shown in Figure 4a and detailed in Table 1. Moreover, the lanthanide-doped samples showed CCT values of 13,546.5, 5770.3, 2028.6, and 3884.2 K, respectively, matching the human eye-friendly application range and photometry implementations [37]. To further assess the quality of the emitted color, the color purity of the emitted color in all the samples was determined through the following equation (Equation (2)) [47]:
c o l o r   p u r i t y = x x s 2 + y y s 2 x d x s 2 + y d y s 2   ×   100
where x and y represent the CIE coordinates of the entire spectrum. xs and ys denote the CIE coordinates of the standard illuminants of white light; and xd and yd stand for the CIE coordinates of the dominant wavelength. Notably, Y-BTC and Tb@Y-BTC exhibited the highest color purity values (see Table 1). However, in some doped samples, the primary transitions were influenced by additional spectral features, affecting color purity. For instance, Eu@Y-BTC presented emission contributions not only from Eu3+ ions but also from the BTC3− ligand, reducing the purity of the red emission. Similarly, in co-doped samples, the red emission from Eu3+ was affected by the intense green emission from Tb3+ ions, which impacted the overall emission profile and the perceived color.
For an in-depth evaluation of the luminescence efficiency of the Eu-containing compounds, the intrinsic quantum yields (QEu) were calculated (see Table 2). This parameter, along with the efficiency of sensitization, determines the overall luminescence quantum yield (QY). Assuming that non-radiative and radiative processes are essentially involved in the depopulation of the 5D0 state, the QEu can be expressed as follows:
Q E u = k r a d k r a d + k n r a d
Thus, QEu of the luminescence expresses how well the radiative processes (krad) compete with the non-radiative processes (knrad) described by Equation (3). In general, non-radiative contributions include back-energy transfer to the sensitizer, electron transfer quenching, and quenching by matrix vibrations. Additionally, vibrations commonly found in organic molecules (C–H, O–H, N–H) can contribute to knrad [48]. The radiative contribution krad can be estimated from the equation:
k r a d = 1 τ r a d
The radiative lifetime τrad can be approximated for Eu(III) by the Equation (5) [49]:
k r a d = A M D , 0 . n 3 . I t o t I M D
Here, AMD,0 is the spontaneous emission probability of the magnetic dipole 5D07F1 transition (14.65 s−1), n is the refractive index (being 1.5 for solids), Itot is the total integrated emission of the 5D07FJ (J = 0–6) transitions and IMD is the integrated emission of the 5D07F1 transition. If the τrad is known, QEu can be calculated using the τobs. Based on Equations (3) and (5), QEu can be calculated as:
Q E u = τ o b s τ r a d
Additionally, by knowing the τobs and τrad it is possible to determine the overall rate of non-radiative deactivation (knrad). Hence, the radiative lifetime is an important parameter for the photophysical description of lanthanide luminescence. The photophysical characterization for the Eu-doped MOFs reported here is summarized in Table 2. According to the PL parameters exhibited in Table 2, it is possible to highlight the promising emission properties of high europium quantum yield and lifetime values in Eu1.25Tb3.75@Y-BTC.

4. Sensing Assays

MOFs have been widely employed in various implementations, including powders, thin films, and composites, for sensing applications. They have demonstrated effectiveness in detecting organic solvents [50], VOCs [51], and metal ions [52]. Additionally, the detection of hazardous analytes, such as explosive molecules [53], is a growing area of interest for MOF-based sensors.
Ln-MOFs, particularly those containing Eu3+ and Tb3+, have gained attention due to their hypersensitive transitions, which provide PL properties that are highly responsive to changes in the chemical and physical environment [54,55]. This feature is relevant for sensor design, either in powder form or anchored into solid substrates. The dual-emission capability of Eu3+ and Tb3+ ions presents a unique opportunity for chemosensing performance.
The chemosensing performance of Eu1.25Tb3.75@Y-BTC was investigated by measuring the PL in the presence of three protic solvents (water, methanol, and ethanol) and six aprotic solvents (DMF, chloroform, toluene, 1,3,5-TMB, and acetone). Upon excitation at 280 nm, the typical Eu3+ and Tb3+transitions are observed in the emission spectra (Figure 5), the 5D07F2 (Eu3+) and 5D47F5 (Tb3⁺) transitions being the most intense.
Comparing the emission profiles of the solid-state samples (black trace) with those of the VOC-exposed (VOC@Eu1.25Tb3.75@Y-BTC), significant variations were observed. Notably, there are certain VOCs, such as ACN, DMF, chloroform, and MeOH, that enhance the 5D07F2 transition intensity by 274, 143, 15, and 6.5%, respectively. In contrast, EtOH, toluene, 1,3,5-TMB, and acetone produce a significant quenching effect, which is detectable to the “naked eye”.
The quenching efficiency (QE) of VOCs was calculated from Equation (7) [56]:
Q E % = I 0 I / I   100
where I0 and I represent the emission intensity values in the absence and presence of the VOC, respectively. The calculated QE values for water, EtOH, toluene, acetone, and 1,3,5-TMB were 89, 94, 99.6, 99.7, and 99.8%.
These findings demonstrate the potential of Eu1.25Tb3.75@Y-BTC as a highly sensitive and selective luminescent sensor for VOC detection, offering promising applications in environmental monitoring and industrial settings.
Understanding the intricate interplay between the host framework and guest VOC molecules is essential for optimizing the sensing performance and expanding the potential applications of luminescent MOF-based sensors. To achieve a comprehensive understanding of the PL behavior, it is essential to analyze additional PL parameters such as krad, knrad, and kexp (kexp = krad + knrad) constants, QEu and τobs (Figures S4 and S5 and Table S1). Figure 6 presents the PL parameters of VOC@Eu1.25Tb3.75@Y-BTC systems, offering insights into the interaction dynamics between the analytes and lanthanide ions.
Analyte–lanthanide ion interactions may be inferred from the determination of energy transfer efficiency within the frame of Förster’s dipole–dipole mechanism that supports the quenching effects [57]. In this context, Equation (1) is also useful to estimate the efficiency of transfer between the donor (Eu or Tb) and the acceptor (VOCs) as follows:
η E u V O C = 1 τ V O C τ s
η T b V O C = 1 τ V O C τ s
The calculated ηEu→VOC were 57.7, 96.9, 97.28 and 98.4% in water, toluene, acetone and 1,3,5-TMB. The corresponding ηTb→VOC values were 99.45, 99,4, 99.8, 99.4, and 99.1% for EtOH, water, toluene, acetone, and 1,3,5-TMB. These values indicate a stronger sensitization of Tb3+ with respect to Eu3+ ions in the presence of the mentioned VOCs.
The quenching mechanism mediated by coupling vibrations is based on the capability of certain atomic groups to dissipate part of the lanthanide energy. This effect can be quantitatively assessed by the so-called “quantum numbers” [49], which represent the number of times such vibrational stretching matches up with the 4f electronic transition. In the present study, the energy of the Eu3+ hypersensitive 5D07F2 transition (16,313 cm−1) corresponds to approximately 4.4 times the energy of -OH (3650 cm−1), 6 times that of -CH (2960 cm−1), and 9.5 times that of -C=O (1680 cm−1) [58].
Similarly, the energy of the 5D47F5 transition of Tb3+ (18,382 cm−1) corresponds to approximately 5, 6.2, and 11 times the vibrational energies of -OH, -CH, and -C=O groups, respectively. Since the lower the quantum number is, the higher the quenching efficiency, hydroxyl (-OH) groups are more efficient in luminescence attenuation compared to other organic groups. Interestingly, this fact aligns more closely with the calculated energy transfer values for ηTb→VOC than for ηEu→VOC, highlighting a differential response of the lanthanide ions to various molecular environments.
Comparing the knrad observed in the Eu1.25Tb3.75@Y-BTC solid, the increased values in the presence of toluene, 1,3,5-TMB, and acetone suggest that the quenching of Eu3+ emission predominantly occurs due to -C=O and -CH groups rather than -OH groups, via non-radiative pathways. Due to the porous nature of the MOF-76 structure [59], these results imply a size–analyte-dependence quenching mechanism facilitated by interactions with lanthanide centers through 6.6 × 6.6 Å2 1D-channels. A similar phenomenon has been previously reported in Eu-BTC frameworks for agrochemical detection [60].
Also, DMF, water, and MeOH, which are commonly coordinated to lanthanide ions in MOF-76, further reinforce the hypothesis that analyte–MOF interactions enhance the sensitization process [52].
Additionally, the solvent-dependent luminescence of Eu3⁺ and Tb3⁺ was analyzed by evaluating competitive pathways between energy transfer and photoinduced electron transfer (PET). Frontier molecular orbital energies (HOMO–LUMO) of the ligand triplet excited state (T1) were calculated in nine analyte-solvents (Table 3).
The values in Table 3 were compared to the excited-state energies of Eu3⁺ (5D0: ~2.0 eV) and Tb3⁺ (5D4: ~2.5–2.7 eV), as well as their redox potentials. The experimental trends— enhanced Eu3⁺ emission in ACN, DMF, methanol, and chloroform versus quenching in acetone, ethanol, water, toluene, and 1,3,5-TMB—are explained below and extended to Tb3⁺ ions.
In ACN, DMF, and methanol, the ligand’s HOMOα (−4.85 to −4.78 eV) lies below the Eu3⁺ excited-state reduction potential (−4.7 eV), suppressing PET (Table 4). The large T1 energy gap (6.2–6.3 eV) exceeds both Eu35D0 and Tb35D4 energies, enabling efficient Dexter-type energy transfer. For Tb3⁺, this alignment suggests strong sensitized emission, assuming minimal non-radiative decay. Chloroform shows distinct behavior; its HOMOα (−3.05 eV) aligns with Eu3⁺ redox potential, but the high LUMOα (2.99 eV) elevates T1 energy (6.04 eV), favoring PET for both Eu3⁺ and Tb3⁺.
In acetone and ethanol, HOMOα (−4.61 to −4.69 eV) approaches the Eu35D0 reduction potential (−4.7 eV), enabling PET (Table 5). For Tb3⁺, T1 energy (6.2 eV) exceeds 5D4 (~2.5–2.7 eV), but PET may compete if the Tb3⁺ excited-state reduction potential is less negative than that of Eu3⁺. Toluene and 1,3,5-TMB exhibit extreme PET due to high HOMOα (−1.03 to −0.64 eV). For Tb3⁺, PET likely dominates despite sufficient T1 energy (5.2–5.7 eV), as the HOMOα alignment overrides energy transfer. Finally, water suppresses emission via solvent polarity effects (e.g., O-H vibrational quenching), destabilizing the T1 state for both Eu3⁺ and Tb3⁺, despite unfavorable PET thermodynamics.
A comparative analysis between Eu3⁺ and Tb3⁺ reveals the following key findings. First, regarding energy transfer efficiency, the T1 energy level (~6 eV) is sufficient to sensitize both ions; however, it is less optimal for Tb3⁺ due to the smaller energy gap between the T1 and 5D4 states compared to the T1 and 5D0 states in Eu3⁺, thereby reducing the efficiency of Förster resonance energy transfer (FRET). Second, in terms of PET competition, the higher 5D4 energy of Tb3⁺ results in a less negative reduction potential. Should the excited-state reduction potential of Tb3⁺ exceed −4.7 eV, PET could compete with other processes in solvent–analytes such as acetone. Third, analyte effects play a significant role: non-polar solvents (e.g., toluene and 1,3,5-TMB) favor PET for both ions, whereas polar solvents like water predominantly induce non-radiative decay pathways, irrespective of the ion species.
These results provide a solid foundation for the development of VOC sensors based on lanthanide coordination polymers, especially for vapor detection devices for air quality monitoring and use in industrial settings.

5. Conclusions

In summary, the photoluminescence analysis of Eu1.25Tb3.75@Y-BTC demonstrates its potential as a highly sensitive and selective sensor for VOC detection. The observed quenching and enhancement effects in the presence of different analytes highlight the significant role of analyte size, functional group interactions, and the porous nature of the MOF-76 framework in modulating the photoluminescence response. The study reveals a size-dependent quenching mechanism, where the analyte diffusion through the 1D channels and the coordination of protic solvents to lanthanide centers significantly influence the sensing performance.
Despite the advances in VOC detection, traditional methods remain limited by high costs, complex instrumentation, and slow response times, emphasizing the need for efficient, recyclable, and easily deployable sensing platforms. The enhanced sensitivity of Tb3⁺ over Eu3⁺ ions towards specific VOCs underscores the potential of dual-emission sensing, offering improved selectivity and real-time monitoring capabilities. Furthermore, the structural stability and recyclability of the Ln@Y-BTC material make it a strong candidate for real-world applications, particularly in air quality monitoring and industrial safety. Understanding the interplay between the host framework and guest VOC molecules is crucial for optimizing sensing performance and broadening the applicability of luminescent MOF-based sensors.
This work not only expands the understanding of lanthanide-based luminescent MOFs for VOC detection but also establishes a foundation for further research aimed at fine-tuning MOF structures for enhanced sensitivity, selectivity, and practical implementation in portable sensing technologies.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/polym17091135/s1. Detailed theoretical calculations such as Optimized Cartesian coordinates of the BTC ligand in the different analytes; Table S1. Quasiharmonic DFT thermodynamic properties; Figures S1–S9 Energy diagram and graphical representation of the ligand frontier orbitals in the analytes; Table S2. Computed absorption energies, excitation state character and transition weight; Figure S10: PXRD patterns of thermal analysis residue compared with the simulated Y₂O₃ pattern; Figure S11. FTIR spectra of Y-BTC and Ln@Y-BTC; Figure S12. Decay profiles of the Ln@Y-BTC solid compounds; Figure S13. Terbium 5D4 decay profiles of VOC@Eu1.25Tb3.75 suspensions; Figure S14. Europium 5D0 decay profiles of VOC@Eu1.25Tb3.75 suspensions; Table S3. Photophysical parameters of VOC@Eu1.25Tb3.75 suspensions; Table S4. Eu and Tb content in co-doped samples determined by ICP-AESP.

Author Contributions

Conceptualization, G.E.G.; Methodology, O.R.R., M.H., M.B. and G.E.G.; Software, H.A.B. and G.E.G.; Validation, M.H., H.A.B., M.B., R.V.D. and G.E.G.; Formal analysis, O.R.R., M.H., M.B., R.V.D. and G.E.G.; Investigation, O.R.R., H.A.B., M.H., M.B., R.V.D. and G.E.G.; Resources, R.V.D. and G.E.G.; Data curation, M.H., M.B., R.V.D. and G.E.G.; Writing—original draft, O.R.R., M.H. and G.E.G.; Writing—review & editing, M.H., M.B., R.V.D. and G.E.G.; Visualization, O.R.R., R.V.D. and G.E.G.; Supervision, M.H., R.V.D. and G.E.G.; Project administration, G.E.G.; Funding acquisition, G.E.G. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by Consejo Nacional de Investigaciones Científicas y Técnicas and PROICO 2320 (UNSL-INTEQUI).

Institutional Review Board Statement

Not applicable.

Data Availability Statement

The original contributions presented in this study are included in the article/supplementary material. Further inquiries can be directed to the corresponding author.

Acknowledgments

This work was supported by the Consejo Nacional de Investigaciones Científicas y Técnicas and PROICO 2320 (UNSL-INTEQUI). O.R. acknowledges support from the Faculty of Chemistry at UNSL for a graduate research scholarship. G.E.G. acknowledges Ghent University’s CESAM Visiting Staff Program. G.E.G., H.A.B. and M.H. are members of CIC-CONICET. This work used computational resources from UNC (Universidad Nacional de Córdoba) Supercómputo (CCAD), which is part of Sistema Nacional de Computación de Alto Desempeño (SNCAD), Argentina.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Kumar, P.; Pournara, A.; Kim, K.-H.; Bansal, V.; Rapti, S.; Manos, M.J. Metal-Organic Frameworks: Challenges and Opportunities for Ion-Exchange/Sorption Applications. Prog. Mater. Sci. 2017, 86, 25–74. [Google Scholar] [CrossRef]
  2. Furukawa, H.; Cordova, K.E.; O’Keeffe, M.; Yaghi, O.M. The Chemistry and Applications of Metal-Organic Frameworks. Science 2013, 341, 1230444. [Google Scholar] [CrossRef]
  3. Qian, Q.; Asinger, P.A.; Lee, M.J.; Han, G.; Mizrahi Rodriguez, K.; Lin, S.; Benedetti, F.M.; Wu, A.X.; Chi, W.S.; Smith, Z.P. MOF-Based Membranes for Gas Separations. Chem. Rev. 2020, 120, 8161–8266. [Google Scholar] [CrossRef]
  4. Bavykina, A.; Kolobov, N.; Khan, I.S.; Bau, J.A.; Ramirez, A.; Gascon, J. Metal–Organic Frameworks in Heterogeneous Catalysis: Recent Progress, New Trends, and Future Perspectives. Chem. Rev. 2020, 120, 8468–8535. [Google Scholar] [CrossRef] [PubMed]
  5. Stavila, V.; Talin, A.A.; Allendorf, M.D. MOF-Based Electronic and Opto-Electronic Devices. Chem. Soc. Rev. 2014, 43, 5994–6010. [Google Scholar] [CrossRef]
  6. Li, H.; Jin, C.; Han, J.; Xi, L.; Song, Z. Tuning Nuclearity of Biradical-Ln Functional Compounds with Single-Molecule Magnet Behavior and Near-Infrared Luminescence. Cryst. Growth Des. 2023, 23, 612–619. [Google Scholar] [CrossRef]
  7. Claudio-Rizo, J.A.; Cano Salazar, L.F.; Flores-Guia, T.E.; Cabrera-Munguia, D.A. Estructuras Metal-Orgánicas (MOFs) Nanoestructuradas Para La Liberación Controlada de Fármacos. Mundo Nano. Rev. Interdiscip. Nanociencias Nanotecnología 2020, 14, 1e–29e. [Google Scholar] [CrossRef]
  8. Gomez, G.E.; Hamer, M.; Regiart, M.D.; Tortella, G.R.; Seabra, A.B.; Soler Illia, G.J.A.A.; Fernández-Baldo, M.A. Advances in Nanomaterials and Composites Based on Mesoporous Materials as Antimicrobial Agents: Relevant Applications in Human Health. Antibiotics 2024, 13, 173. [Google Scholar] [CrossRef]
  9. Zhou, H.-C.; Long, J.R.; Yaghi, O.M. Introduction to Metal–Organic Frameworks. Chem. Rev. 2012, 112, 673–674. [Google Scholar] [CrossRef]
  10. Shi, X.; Shan, Y.; Du, M.; Pang, H. Synthesis and Application of Metal-Organic Framework Films. Coord. Chem. Rev. 2021, 444, 214060. [Google Scholar] [CrossRef]
  11. Zhang, Y.; Chang, C.-H. Metal–Organic Framework Thin Films: Fabrication, Modification, and Patterning. Processes 2020, 8, 377. [Google Scholar] [CrossRef]
  12. Zhang, R.; Zhu, L.; Yue, B. Luminescent Properties and Recent Progress in Applications of Lanthanide Metal-Organic Frameworks. Chin. Chem. Lett. 2023, 34, 108009. [Google Scholar] [CrossRef]
  13. Song, X.-Z.; Song, S.-Y.; Zhang, H.-J. Luminescent Lanthanide Metal–Organic Frameworks; Springer: Berlin/Heidelberg, Germany, 2014; pp. 109–144. [Google Scholar]
  14. Gomez, G.E.; Roncaroli, F. Photofunctional Metal-Organic Framework Thin Films for Sensing, Catalysis and Device Fabrication. Inorganica Chim. Acta 2020, 513, 119926. [Google Scholar] [CrossRef]
  15. Herrera, F.C.; Caraballo, R.M.; Soler Illia, G.J.A.A.; Gomez, G.E.; Hamer, M. Sunlight-Driven Photocatalysis for a Set of 3D Metal–Porphyrin Frameworks Based on a Planar Tetracarboxylic Ligand and Lanthanide Ions. ACS Omega 2023, 8, 46777–46785. [Google Scholar] [CrossRef] [PubMed]
  16. SeethaLekshmi, S.; Ramya, A.R.; Reddy, M.L.P.; Varughese, S. Lanthanide Complex-Derived White-Light Emitting Solids: A Survey on Design Strategies. J. Photochem. Photobiol. C Photochem. Rev. 2017, 33, 109–131. [Google Scholar] [CrossRef]
  17. Rosi, N.L.; Kim, J.; Eddaoudi, M.; Chen, B.; O’Keeffe, M.; Yaghi, O.M. Rod Packings and Metal−Organic Frameworks Constructed from Rod-Shaped Secondary Building Units. J. Am. Chem. Soc. 2005, 127, 1504–1518. [Google Scholar] [CrossRef]
  18. Duan, T.-W.; Yan, B. Hybrids Based on Lanthanide Ions Activated Yttrium Metal–Organic Frameworks: Functional Assembly, Polymer Film Preparation and Luminescence Tuning. J. Mater. Chem. C 2014, 2, 5098–5104. [Google Scholar] [CrossRef]
  19. Jiang, C.; Wang, X.; Ouyang, Y.; Lu, K.; Jiang, W.; Xu, H.; Wei, X.; Wang, Z.; Dai, F.; Sun, D. Recent Advances in Metal–Organic Frameworks for Gas Adsorption/Separation. Nanoscale Adv. 2022, 4, 2077–2089. [Google Scholar] [CrossRef]
  20. Brunckova, H.; Mudra, E.; Shepa, I. Recent Advances in Lanthanide Metal–Organic Framework Thin Films Based on Eu, Tb, Gd: Preparation and Application as Luminescent Sensors and Light-Emitting Devices. Inorganics 2023, 11, 376. [Google Scholar] [CrossRef]
  21. Shen, Y.; Tissot, A.; Serre, C. Recent Progress on MOF-Based Optical Sensors for VOC Sensing. Chem. Sci. 2022, 13, 13978–14007. [Google Scholar] [CrossRef]
  22. Kau, N.; Jindal, G.; Kaur, R.; Rana, S. Progress in Development of Metal Organic Frameworks for Electrochemical Sensing of Volatile Organic Compounds. Results Chem. 2022, 4, 100678. [Google Scholar] [CrossRef]
  23. Afifa; Arshad, K.; Hussain, N.; Ashraf, M.H.; Saleem, M.Z. Air Pollution and Climate Change as Grand Challenges to Sustainability. Sci. Total Environ. 2024, 928, 172370. [Google Scholar] [CrossRef]
  24. Jin, X.; Wu, Y.; Santhamoorthy, M.; Nhi Le, T.T.; Le, V.T.; Yuan, Y.; Xia, C. Volatile Organic Compounds in Water Matrices: Recent Progress, Challenges, and Perspective. Chemosphere 2022, 308, 136182. [Google Scholar] [CrossRef] [PubMed]
  25. Paolin, E.; Strlič, M. Volatile Organic Compounds (VOCs) in Heritage Environments and Their Analysis: A Review. Appl. Sci. 2024, 14, 4620. [Google Scholar] [CrossRef]
  26. Epping, R.; Koch, M. On-Site Detection of Volatile Organic Compounds (VOCs). Molecules 2023, 28, 1598. [Google Scholar] [CrossRef] [PubMed]
  27. Han, B.; Rupam, T.H.; Chakraborty, A.; Saha, B.B. A Comprehensive Review on VOCs Sensing Using Different Functional Materials: Mechanisms, Modifications, Challenges and Opportunities. Renew. Sustain. Energy Rev. 2024, 196, 114365. [Google Scholar] [CrossRef]
  28. Okur, S.; Hashem, T.; Bogdanova, E.; Hodapp, P.; Heinke, L.; Bräse, S.; Wöll, C. Optimized Detection of Volatile Organic Compounds Utilizing Durable and Selective Arrays of Tailored UiO-66-X SURMOF Sensors. ACS Sens. 2024, 9, 622–630. [Google Scholar] [CrossRef]
  29. Cao, Y.; Fu, M.; Fan, S.; Gao, C.; Ma, Z.; Hou, D. Hydrophobic MOF/PDMS-Based QCM Sensors for VOCs Identification and Quantitative Detection in High-Humidity Environments. ACS Appl. Mater. Interfaces 2024, 16, 7721–7731. [Google Scholar] [CrossRef]
  30. Peng, X.; Wu, X.; Zhang, M.; Yuan, H. Metal–Organic Framework Coated Devices for Gas Sensing. ACS Sens. 2023, 8, 2471–2492. [Google Scholar] [CrossRef]
  31. Chen, J.; Zhang, R.; Guo, S.; Pan, Y.; Nezamzadeh-Ejhieh, A.; Lan, Q. Metal-Organic Frameworks (MOFs): A Review of Volatile Organic Compounds (VOCs) Detection. Talanta 2025, 286, 127498. [Google Scholar] [CrossRef]
  32. Liu, X.; Lin, K.; Chang, J. Modulation of Hydroxyapatite Crystals Formed from α-Tricalcium Phosphate by Surfactant-Free Hydrothermal Exchange. CrystEngComm 2011, 13, 1959–1965. [Google Scholar] [CrossRef]
  33. Jiang, H.-L.; Tsumori, N.; Xu, Q. A Series of (6,6)-Connected Porous Lanthanide−Organic Framework Enantiomers with High Thermostability and Exposed Metal Sites: Scalable Syntheses, Structures, and Sorption Properties. Inorg. Chem. 2010, 49, 10001–10006. [Google Scholar] [CrossRef]
  34. Frisch, M.J.; Trucks, G.W.; Schlegel, H.B.; Scuseria, G.E.; Robb, M.a.; Cheeseman, J.R.; Scalmani, G.; Barone, V.; Petersson, G.a.; Nakatsuji, H.; et al. G16_C01 2016, Gaussian 16, Revision C.01; Gaussian, Inc.: Wallingford, CT, USA, 2016. [Google Scholar]
  35. Chai, J.-D.; Head-Gordon, M. Long-Range Corrected Hybrid Density Functionals with Damped Atom–Atom Dispersion Corrections. Phys. Chem. Chem. Phys. 2008, 10, 6615. [Google Scholar] [CrossRef] [PubMed]
  36. Mardirossian, N.; Head-Gordon, M. Thirty Years of Density Functional Theory in Computational Chemistry: An Overview and Extensive Assessment of 200 Density Functionals. Mol. Phys. 2017, 115, 2315–2372. [Google Scholar] [CrossRef]
  37. Krishnan, R.; Binkley, J.S.; Seeger, R.; Pople, J.A. Self-Consistent Molecular Orbital Methods. XX. A Basis Set for Correlated Wave Functions. J. Chem. Phys. 1980, 72, 650–654. [Google Scholar] [CrossRef]
  38. Cramer, C.J. Essentials of Computational Chemistry: Theories and Models; Wiley: Hoboken, NJ, USA, 2004. [Google Scholar]
  39. Dreuw, A.; Head-Gordon, M. Single-Reference Ab Initio Methods for the Calculation of Excited States of Large Molecules. Chem. Rev. 2005, 105, 4009–4037. [Google Scholar] [CrossRef]
  40. Tomasi, J.; Mennucci, B.; Cammi, R. Quantum Mechanical Continuum Solvation Models. Chem. Rev. 2005, 105, 2999–3094. [Google Scholar] [CrossRef]
  41. Casida, M.E. Time-Dependent Density Functional Response Theory for Molecules. Recent Adv. Density Funct. Methods 1995, 155–192. [Google Scholar] [CrossRef]
  42. Cui, Y.; Yue, Y.; Qian, G.; Chen, B. Luminescent Functional Metal–Organic Frameworks. Chem. Rev. 2012, 112, 1126–1162. [Google Scholar] [CrossRef]
  43. Cui, Y.; Xu, H.; Yue, Y.; Guo, Z.; Yu, J.; Chen, Z.; Gao, J.; Yang, Y.; Qian, G.; Chen, B. A Luminescent Mixed-Lanthanide Metal–Organic Framework Thermometer. J. Am. Chem. Soc. 2012, 134, 3979–3982. [Google Scholar] [CrossRef]
  44. Li, H.; Jing, P.; Lu, J.; Xi, L.; Wang, Q.; Ding, L.; Wang, W.-M.; Song, Z. Multifunctional Properties of {CuII2LnIII2} Systems Involving Nitrogen-Rich Nitronyl Nitroxide: Single-Molecule Magnet Behavior, Luminescence, Magnetocaloric Effects and Heat Capacity. Dalt. Trans. 2021, 50, 2854–2863. [Google Scholar] [CrossRef] [PubMed]
  45. Chen, D.; Sedykh, A.E.; Gomez, G.E.; Neumeier, B.L.; Santos, J.C.C.; Gvilava, V.; Maile, R.; Feldmann, C.; WÖll, C.; Janiak, C.; et al. SURMOF Devices Based on Heteroepitaxial Architectures with White-Light Emission and Luminescent Thermal-Dependent Performance. Adv. Mater. Interfaces 2020, 7, 2000929. [Google Scholar] [CrossRef]
  46. Munirathnam, K.; Dillip, G.R.; Ramesh, B.; Joo, S.W.; Prasad Raju, B.D. Synthesis, Photoluminescence and Thermoluminescence Properties of LiNa3P2O7:Tb3+ Green Emitting Phosphor. J. Phys. Chem. Solids 2015, 86, 170–176. [Google Scholar] [CrossRef]
  47. Available online: https://sciapps.sci-sim.com/ (accessed on 2 December 2024).
  48. Gomez, G.E.; Bernini, M.C.; Brusau, E.V.; Narda, G.E.; Vega, D.; Kaczmarek, A.M.; Van Deun, R.; Nazzarro, M. Layered Exfoliable Crystalline Materials Based on Sm-, Eu- and Eu/Gd-2-Phenylsuccinate Frameworks. Crystal Structure, Topology and Luminescence Properties. Dalt. Trans. 2015, 44, 3417–3429. [Google Scholar] [CrossRef] [PubMed]
  49. Chauvin, A.; Gumy, F.; Imbert, D.; Bünzli, J.G. Europium and Terbium Tris (Dipicolinates) as Secondary Standards for Quantum Yield Determination. Spectrosc. Lett. 2004, 37, 517–532. [Google Scholar] [CrossRef]
  50. Chen, B.; Yang, Y.; Zapata, F.; Lin, G.; Qian, G.; Lobkovsky, E.B. Luminescent Open Metal Sites within a Metal–Organic Framework for Sensing Small Molecules. Adv. Mater. 2007, 19, 1693–1696. [Google Scholar] [CrossRef]
  51. Xu, H.; Rao, X.; Gao, J.; Yu, J.; Wang, Z.; Dou, Z.; Cui, Y.; Yang, Y.; Chen, B.; Qian, G. A Luminescent Nanoscale Metal–Organic Framework with Controllable Morphologies for Spore Detection. Chem. Commun. 2012, 48, 7377. [Google Scholar] [CrossRef]
  52. Chen, B.; Wang, L.; Zapata, F.; Qian, G.; Lobkovsky, E.B. A Luminescent Microporous Metal−Organic Framework for the Recognition and Sensing of Anions. J. Am. Chem. Soc. 2008, 130, 6718–6719. [Google Scholar] [CrossRef]
  53. Asad, M.; Anwar, M.I.; Miao, B.; Abbas, A.; Majeed, S.; Mir, I.A.; Rabbani, M.S.; Hussain, S.; Xu, S.; Al-Tahan, M.A.; et al. Recent Advances in Luminescent Metal-Organic Frameworks (L-MOFs) as Sustainable Materials for Sensing of Potentially Toxic Environmental Ubiquitous Explosive Contaminants. Sustain. Mater. Technol. 2024, 42, e01155. [Google Scholar] [CrossRef]
  54. Trannoy, V.; Carneiro Neto, A.N.; Brites, C.D.S.; Carlos, L.D.; Serier-Brault, H. Engineering of Mixed Eu3+ /Tb3+ Metal-Organic Frameworks Luminescent Thermometers with Tunable Sensitivity. Adv. Opt. Mater. 2021, 9, 2001938. [Google Scholar] [CrossRef]
  55. Wang, X.; Ma, T.; Ma, J.-G.; Cheng, P. Integration of Devices Based on Metal–Organic Frameworks: A Promising Platform for Chemical Sensing. Coord. Chem. Rev. 2024, 518, 216067. [Google Scholar] [CrossRef]
  56. Godoy, A.A.; Gomez, G.E.; Kaczmarek, A.M.; Van Deun, R.; Furlong, O.J.; Gándara, F.; Monge, M.A.; Bernini, M.C.; Narda, G.E. Sensing Properties, Energy Transfer Mechanism and Tuneable Particle Size Processing of Luminescent Two-Dimensional Rare Earth Coordination Networks. J. Mater. Chem. C 2017, 5, 12409–12421. [Google Scholar] [CrossRef]
  57. Piguet, C.; Buenzli, J.C.G.; Bernardinelli, G.; Hopfgartner, G.; Williams, A.F. Self-Assembly and Photophysical Properties of Lanthanide Dinuclear Triple-Helical Complexes. J. Am. Chem. Soc. 1993, 115, 8197–8206. [Google Scholar] [CrossRef]
  58. Hybrid Materials: Synthesis, Characterization, and Applications; Kickelbick, G., Ed.; Wiley-VCH Verlag GmbH & Co.: Weinheim, Germany, 2007. [Google Scholar]
  59. Meng, D.; Zhao, T.; Busko, D.; Cosgun Ergene, A.; Richards, B.S.; Howard, I.A. Tb and Eu in MOF-76: Elucidating the Mechanisms Responsible for the Divergent Excellent and Poor Photoluminescence Quantum Yields. Adv. Opt. Mater. 2024, 12, 2300867. [Google Scholar] [CrossRef]
  60. Gomez, G.E.; dos Santos Afonso, M.; Baldoni, H.A.; Roncaroli, F.; Soler-Illia, G.J.A.A. Luminescent Lanthanide Metal Organic Frameworks as Chemosensing Platforms towards Agrochemicals and Cations. Sensors 2019, 19, 1260. [Google Scholar] [CrossRef]
Figure 1. (a) Crystal of Y-BTC. (b) Lanthanide environment of one metallic center. (c) Projection on the ab and ac planes of Y-BTC (color code: brown: carbon; red: oxygen; white: hydrogen; blue: yttrium). (d) Topological view of the MOF-76 structure.
Figure 1. (a) Crystal of Y-BTC. (b) Lanthanide environment of one metallic center. (c) Projection on the ab and ac planes of Y-BTC (color code: brown: carbon; red: oxygen; white: hydrogen; blue: yttrium). (d) Topological view of the MOF-76 structure.
Polymers 17 01135 g001
Figure 2. (a) PXRD pattern of Ln@Y-BTC compounds compared to the simulated one from the Y-BTC structure [33]. (b) TGA/DSC plots from the Eu@Y-BTC compound.
Figure 2. (a) PXRD pattern of Ln@Y-BTC compounds compared to the simulated one from the Y-BTC structure [33]. (b) TGA/DSC plots from the Eu@Y-BTC compound.
Polymers 17 01135 g002
Figure 3. (a) Excitation spectra of Y-BTC and Ln@Y-BTC samples; emission profiles from (b) Y-BTC, (c) Eu@Y-BTC, (d) Tb@Y-BTC, (e) Eu2.5Tb2.5@Y-BTC, and (f) Eu1.25Tb3.75@Y-BTC upon excitation at 280 nm.
Figure 3. (a) Excitation spectra of Y-BTC and Ln@Y-BTC samples; emission profiles from (b) Y-BTC, (c) Eu@Y-BTC, (d) Tb@Y-BTC, (e) Eu2.5Tb2.5@Y-BTC, and (f) Eu1.25Tb3.75@Y-BTC upon excitation at 280 nm.
Polymers 17 01135 g003
Figure 4. (a) CIE diagram showing the color coordinates; (b) the proposed Jablonski diagram for the compounds reported herein.
Figure 4. (a) CIE diagram showing the color coordinates; (b) the proposed Jablonski diagram for the compounds reported herein.
Polymers 17 01135 g004
Figure 5. (a) PL spectra of VOC@Eu1.25Tb3.75@Y-BTC suspensions recorded at room temperature (λexc = 280 nm); (b) inset into the 5D07F2 signal and (c) their color effects under UV light exposure.
Figure 5. (a) PL spectra of VOC@Eu1.25Tb3.75@Y-BTC suspensions recorded at room temperature (λexc = 280 nm); (b) inset into the 5D07F2 signal and (c) their color effects under UV light exposure.
Polymers 17 01135 g005
Figure 6. (a) QEu (%) and lifetime values; (b) krad and knrad constants of VOC@Eu1.25Tb3.75@Y-BTC systems.
Figure 6. (a) QEu (%) and lifetime values; (b) krad and knrad constants of VOC@Eu1.25Tb3.75@Y-BTC systems.
Polymers 17 01135 g006
Table 1. CIE chromaticity, CCT values, and calculated color purity from all the compounds.
Table 1. CIE chromaticity, CCT values, and calculated color purity from all the compounds.
SampleCIE Chromaticity from the Entire SpectrumCCT (K)CIE Chromaticity from the Dominant WavelengthColor Purity (%)
xyxdyd
Y-BTC0.1440.071937.40.170.00689.97
Eu@Y-BTC0.2810.2513,546.50.680.3121.28
Tb@Y-BTC0.3190.5455770.30.260.7355.75
Eu2.5Tb2.5@Y-BTC0.4740.3542028.60.680.3145.11
Eu1.25Tb3.75@Y-BTC0.3840.3733884.20.680.3125.76
Table 2. Photophysical parameters of the europium-based MOFs.
Table 2. Photophysical parameters of the europium-based MOFs.
CompoundItot/IMDτrad/mskrad/s−1kexp/s−1knrad/s−1τobs/ms QEu (%)
Eu@Y-BTC14.521.39718.13571.422853.320.2820.1
Eu2.5Tb2.5@Y-BTC9.582.11474.053508.773034.710.28513.51
Eu1.25Tb3.75@Y-BTC10.91.85539.332564.12024.760.3921.03
Table 3. Energy levels (eV) of the ligand in various analytes. This table presents the HOMO and LUMO energies for both α and β electrons, along with their respective energy gaps (ΔHLα/β).
Table 3. Energy levels (eV) of the ligand in various analytes. This table presents the HOMO and LUMO energies for both α and β electrons, along with their respective energy gaps (ΔHLα/β).
AnalyteHOMOaLUMOaHOMObLUMObΔHLaΔHLbΔHLtot.
ACN−4.85331.4235−8.0813−2.04376.27716.03922.8091
DMF−4.81731.3894−8.1367−2.10776.20646.03032.7096
Chloroform−3.04792.988−6.3256−0.3386.03755.9882.7096
Methanol−4.78041.425−8.0983−2.07166.2066.02942.7093
Water−5.00721.2801−8.2273−2.18546.28766.04332.8214
Ethanol−4.68931.515−8.0041−1.98336.20466.02272.7063
Toluene−1.03254.6158−4.27541.69075.6485.96652.7232
Acetone−4.60731.5966−7.9209−1.90156.2045.92042.706
1,3,5-TMB−0.63534.5709−4.03022.25215.20676.28262.8865
Table 4. Group A solvents: enhanced emission (energy transfer dominates).
Table 4. Group A solvents: enhanced emission (energy transfer dominates).
AnalyteHOMOα (eV)LUMOα (eV)T1 Energy (eV)Relevance to Tb3+
ACN−4.851.426.27“T1 >> Tb35D4 (~2.5–2.7 eV); efficient energy transfer”
DMF−4.821.396.21“Similar to ACN”
Methanol−4.781.436.21“enough T1 energy for Tb3⁺ sensitization”
Chloroform−3.052.996.04“High T1 ensures PET despite moderate HOMOα”
Table 5. Group B solvents: quenched emission (PET or analyte effects).
Table 5. Group B solvents: quenched emission (PET or analyte effects).
SolventHOMOα (eV)LUMOα (eV)T1 Energy (eV)Primary Quenching MechanismRelevance to Tb3+
Acetone−4.611.66.21“PET (HOMOα ≈ −4.7 eV)”“PET competes with energy transfer (T1 > 5D4)”
Ethanol−4.691.526.21“Marginal PET competition”“Similar to acetone”
Water−5.011.286.29“Solvent-induced non-radiative decay”“Polarity disrupts T1 state for both ions”
Toluene−1.034.625.65“Strong PET (high HOMOα)”“PET dominates despite T1 > 5D4
1,3,5-TMB−0.644.575.21“Extreme PET (very high HOMOα)”“PET overrides energy transfer”
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

Rosas Rivas, O.; Hamer, M.; Baldoni, H.A.; Boone, M.; Van Deun, R.; Gomez, G.E. Engineering Photoluminescence of Lanthanide Doped Yttrium-MOF-76 for Volatile Organic Compound Sensing. Polymers 2025, 17, 1135. https://doi.org/10.3390/polym17091135

AMA Style

Rosas Rivas O, Hamer M, Baldoni HA, Boone M, Van Deun R, Gomez GE. Engineering Photoluminescence of Lanthanide Doped Yttrium-MOF-76 for Volatile Organic Compound Sensing. Polymers. 2025; 17(9):1135. https://doi.org/10.3390/polym17091135

Chicago/Turabian Style

Rosas Rivas, Oswaldo, Mariana Hamer, Héctor A. Baldoni, Maya Boone, Rik Van Deun, and Germán E. Gomez. 2025. "Engineering Photoluminescence of Lanthanide Doped Yttrium-MOF-76 for Volatile Organic Compound Sensing" Polymers 17, no. 9: 1135. https://doi.org/10.3390/polym17091135

APA Style

Rosas Rivas, O., Hamer, M., Baldoni, H. A., Boone, M., Van Deun, R., & Gomez, G. E. (2025). Engineering Photoluminescence of Lanthanide Doped Yttrium-MOF-76 for Volatile Organic Compound Sensing. Polymers, 17(9), 1135. https://doi.org/10.3390/polym17091135

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop