Palak
Chaudhary
a,
Sahil
Thakur
ab,
Raghubir
Singh
*b and
Varinder
Kaur
*a
aDepartment of Chemistry, Panjab University, Sector-14, Chandigarh-160014, India. E-mail: var_ka04@yahoo.co.in; var_ka04@pu.ac.in
bDepartment of Chemistry, DAV College, Sector 10, Chandigarh-160011, India. E-mail: raghu_chem2006@yahoo.com; raghubirsingh@davchd.ac.in
First published on 10th July 2025
In this work, the synthesis of luminescent Eu(III)-encapsulated Cu MOF and Zn MOF is reported as efficient materials with tunable light-emitting properties. The MOFs were derived from a salen-type Schiff base with multiple N donors and additional –OH groups for holding Eu(III). The Eu(III)-encapsulated MOFs, Eu–Cu MOF and Eu–Zn MOF hybrids, were characterized using FTIR, TGA, PXRD, FE-SEM, and XPS analyses. The hybrids emitted a wide range of light in the visible and near-IR range. The International Commission on Illumination (CIE) coordinates of Eu–Cu MOF and Eu–Zn MOF hybrids were (0.25, 0.35) and (0.31, 0.29), with correlated colour temperature (CCT) magnitudes of 9896 K and 6719 K, respectively. Additionally, when the outer epoxy resin layer of LEDs was uniformly coated with Eu–Cu MOF and Eu–Zn MOF, green light and near-white light were emitted, respectively, at a voltage of 3 V. Thus, this work provides a new synthetic approach for the development of white light-emitting host–guest materials.
Transition metal coordination frameworks exhibit photoluminescence in a specific range of the visible spectrum; however, the introduction of a guest chromophore can make these materials multi-wavelength emitters.23 Notably, lanthanide-incorporated frameworks show improved white light emission, narrow emission band, significant antenna-generated shift, multiband emission, and excellent luminescence efficiency owing to the antenna effect of ligands.24 Earlier, some aromatic N-heterocycles (including imidazole, pyrrole, pyrimidine, pyridine, and their derivatives) and multifunctional carboxylic acids have been utilized to construct MOFs for light-emitting applications.25 The N-based ligands contribute substantially to the photoluminescence of complexes, driving the excitation spectrum from the ultraviolet to the near-infrared or visible region and providing a high coordination affinity towards metal centres. It is well known that the Zn2+ ions show a great tendency to coordinate with N donor ligands and display an extra edge for inducing photoluminescence properties due to the d10 system.26 Here, we report the synthesis of high-quality crystalline 1D Zn-MOF and Cu-MOF under mild reaction conditions using an N-equipped compartmental ligand. This ligand possesses some interesting chemical properties and binding behaviour due to the presence of the –OH functional group. The –OH group exhibits variable binding modes; it divides the donor set into two compartments and can hold two metal ions in separate compartments or it protrudes outward leaving the
donor set for coordinating a single metal ion in the compartment. Earlier, the inward orientation of the –OH groups has been reported to hold Dy3+ and Cu2+ ions, thus producing dinuclear complexes.27,28 In contrast, the formation of mononuclear complexes has been observed with Mn2+ and Cu2+, indicating the non-involvement of –OH due to its outward orientation.29,30 In another case, the Cu complex was extended to a 1D coordination polymer by the involvement of chlorate ions as bridges.30 The presence of –OH has also been reported to enhance the overall activity of the molecule and facilitate Schiff base cyclization of the ligand in the presence of Fe(II), which induces the oxidative DNA cleavage activity in the molecule.31 In the current work, the –OH acts as an additional active site for holding Eu(III) ions within the channels formed by 1D Zn-MOF and Cu-MOF. The Eu(III)-encapsulated MOFs produced near-white light and greenish light emissions in the Eu–Zn MOF and Eu–Cu MOF hybrids, respectively, demonstrating their potential for application in optical devices. The host–guest combination of hybrids results in reduced molecular vibration, which diminishes non-radiative relaxation and thus improves multiband emission efficiency.32 Moreover, to verify the physical significance of the synthesized material, a prototype was developed by coating Eu–Cu MOF and Eu–Zn MOF on the plastic epoxy resin cover of LEDs. The fabrication of Eu(III)-encapsulated MOFs provides an alternative way for constructing efficient optoelectronic devices.
Yield: 75% (1.12 g); FT-IR (ATR) cm−1: 3371 (broad O–H), 2910 (C–H), 1648 (CN), 1581 (C
Npyridine), 1431 (C–C), 1088 (C–O), 763 (C–H bending). 1H NMR (500 MHz, DMSO-d6): δ (ppm) 3.68 (2H, m, H2b), 3.86 (2H, m, H2a), 4.10 (1H, m, H1), 4.93 (alc OH, broad singlet), 7.46 (2H, m, H7), 7.88 (2H, m, H6), 7.99 (2H, dt, 3J = 5 Hz, H5), 8.36 (2H, s, H3), 8.65 (2H, dq, 3J = 5 Hz, H8); 13C NMR (125 MHz, DMSO-d6): δ (ppm) 64.8 (C2), 69.5 (C1), 120.4 (C5), 124.9 (C7), 136.7 (C6), 149.2 (C8), 154.1 (C3), 162.9 (C4); ESI-MS (m/z): 271.08 [L + 3H]+, 285.09 [L + NH3/ OH˙]+, 181.06 [C9H15N3O]+.
Yield (crystals): 53% (0.15 g), melting point: 222–224 °C; FT-IR (ATR) cm−1: 3259 (broad O–H), 1660 (CN), 1605 (C
Npyridine), 1378 (C–C), 1302 (C–N), 1049 (N–O), 780 (C–H bend), 644 (Cu–N), 568 (Cu–Npyridine). ESI-MS (m/z): 331.05 [M + H]+.
The FT-IR spectrum of the SB ligand shows a broad band at 3371 cm−1 due to O–H stretching, two sharp bands at 1648 cm−1 and 1581 cm−1 due to two types of CN stretching vibrations (azomethinic C
N and C
Npyridine), and a sharp band at 1431 cm−1 due to C–C stretching (ESI,† Fig. S1). The C
N band centred at 1648 cm−1 in the free ligand is shifted to 1660 cm−1 and 1653 cm−1 in Cu MOF and Zn MOF, respectively, due to the formation of the Cu/Zn–N bond. The N–O vibration of nitrate bridges is observed at 1049 cm−1 and 1022 cm−1 in Cu MOF and Zn MOF, respectively.34 An additional band at 644 cm−1 in Cu MOF and 639 cm−1 in Zn MOF confirms the formation of the Cu–N and Zn–N bonds, respectively. The Cu–Npyridine and Zn–Npyridine bond formation is confirmed by the presence of additional bands at 568 cm−1 and 505 cm−1 in Cu MOF and Zn MOF, respectively.
In the 1H NMR, the SB ligand exhibits a singlet at 8.36 ppm for H3, at 4.10 ppm for H1, and two separate signals at 3.68 ppm and 3.86 ppm for H2 protons due to geminal coupling (ESI,† Fig. S2). In Zn MOF, the doublet of quartets of H8 merges into a doublet and shifts from 8.65 to 9.00 ppm, the singlet at 8.36 shifts to 8.85 ppm for H3, the doublet of triplets at 7.99 merges into a doublet and shifts to 8.18 ppm for H5 (ESI,† Fig. S3). Also, 13C NMR of the SB ligand reveals the presence of CN at 154.1 ppm and C
N pyridine at 149.2 ppm (ESI,† Fig. S4). In Zn MOF, C
N shifts to 152.9 ppm from 154.1 ppm (ESI,† Fig. S5). The ESI-MS of ligand SB exhibits a molecular ion peak at 271.08 due to (M + 3H)+ and a fragmentation peak at 285.09 corresponding to [M + NH3/OH˙] (ESI,† Fig. S6). The ESI-MS for Cu MOF and Zn MOF exhibits a peak at 331.05 for [M + H]+ and at 340.45 for [M–OH + Na + H]+, respectively (ESI,† Fig. S7 and S8). The 1H and 13C NMR were not recorded for Cu MOF owing to the very poor solubility and paramagnetic nature of Cu(II). The Cu(II) complexes display broad signals and poor spectral resolution due to the rapid nuclear relaxation of paramagnetic Cu(II) centres.35
The structural parameters and binding behaviour of the 1D MOFs differ from most of the previously reported complexes of this ligand. However, they closely resemble the [Cu(HL)(ClO4)] MOF, where bridges are formed by perchlorate ions.30 The metal centres are six-coordinate in these MOFs with an N4O2 coordination geometry. Here, the Schiff base ligand coordinates with copper and zinc as a tetradentate ligand through two pyridine nitrogen atoms and two imine nitrogen atoms. The elongated axial positions are occupied by nitrate ligands. A single-crystal X-ray diffraction analysis reveals that the Cu MOF crystallized in the P21/n space group of the monoclinic crystal system. In the structure shown in Fig. 1a, the Cu(II) ion is coordinated with the nitrogen atoms from both imine groups (N2 and N3), the nitrogen atoms from the pyridine rings (N1 and N4), and the oxygen atoms of two nitrate ions, resulting in a distorted octahedral geometry. The bond lengths are found to be Cu–N3 (1.997 Å), Cu–N2 (2.004 Å), Cu–N4 (2.079 Å), and Cu–N1 (2.015 Å). The Cu MOF network is further extended by nitrate bridging, with a Cu–O2 bond length of 2.266 Å (Fig. 1b). Similarly, the Zn MOF crystallized in the P1 space group of the triclinic crystal system. Its asymmetric unit comprises two distinct Zn(II) ion centres, Zn1 and Zn2, both exhibiting similar coordination with nitrogen atoms from the imine groups (N2 and N3) and the pyridine nitrogen atoms (N1 and N4) as shown in Fig. 1c. Zn1 is coordinated to two nitrate ions, while Zn2 is coordinated to one nitrate ion and one water molecule, resulting in a distorted octahedral geometry at both centres. Additionally, there are nitrate ions present within the lattice, each of which is hydrogen-bonded to a Zn1 centre through an (O1H---O–N, 2.162 Å) interaction and to two Zn2 centres via (HO12H---O–N, 2.088 Å) interactions. This unique binding arrangement formed a 1D architecture comprising two parallel chains, one featuring Zn1-bonded molecules and the other containing Zn2-bonded molecules. The hydroxyl group of 1,3-diamino-2-propanol does not participate in binding and remains free with outward orientation from the coordination sphere. This may be due to steric hindrance, electronic factors, or lack of proper orientation for effective binding with metal ions, leading the coordination sphere to involve mainly nitrogen atoms and nitrate ions (Fig. 1d). Crystallographic data and parameters used for structure refinement are given in ESI† Tables S1 and S2 provides information about specific measurements of bond lengths and bond angles.
![]() | ||
Fig. 1 (a) Asymmetric unit of a Cu MOF, (b) side view of the Cu MOF framework, (c) asymmetric unit of a Zn MOF, and (d) top-down view of the Zn MOF framework. |
The presence of a free hydroxyl group with an outward orientation provides an interactive site to hold the metal ions within the 1D layers of MOFs. Thus, Eu(III) ions are embedded on the surface of these MOFs, which form a coating on the surface of the MOFs. The FT-IR spectra of Eu–Cu MOF and Eu–Zn MOF resemble the parent MOFs (Fig. 2a and b). The spectrum of Eu–Cu MOF shows a broad band at 3340 cm−1 for O–H stretching, which is probably due to solvated Eu(III) ions. A shifting in the bands for CN from 1660 cm−1 to 1635 cm−1, C
Npyridine from 1605 cm−1 to 1599 cm−1, N–O from 1049 cm−1 to 1025 cm−1,34 Cu–N from 644 cm−1 to 631 cm−1, and Cu–Npyridine from 568 cm−1 to 504 cm−1 is observed, suggesting the dispersion of Eu(III) ions over the surface of the MOF layers through π-interactions. A similar trend was observed in the case of Eu–Zn MOF, where the broadening of the O–H band is found along with the shifting of bands (C
N from 1653 cm−1 to 1650 cm−1, C
Npyridine from 1596 cm−1 to 1588 cm−1, N–O peak from 1022 cm−1 to 1014 cm−1, Zn–N from 639 cm−1 to 631 cm−1, and Zn–Npyridine from 505 cm−1 to 501 cm−1). Both Eu–Cu MOF and Eu–Zn MOF retain their characteristic vibrational bands in their FT-IR (ATR) spectra, which account for the stability and integrity of the host frameworks after the infusion of Eu(III) ions in the layers.
The thermal behaviour of Cu MOF, Eu–Cu MOF, Zn MOF, and Eu–Zn MOF was examined by thermogravimetry, and the respective curves are presented in Fig. 3. In Cu MOF, the initial mass loss of 2.80% up to 230 °C is attributed to the escape of solvent and trapped water molecules. When the temperature is increased to 230 °C, a second weight loss of 68.02% is observed due to the breakdown of Cu MOF. The third mass loss of 7% that occurs after the temperature is raised above 235 °C is related to the formation of CuO and Eu2O3, resulting from the disintegration of the complete skeleton.36 Similarly, a large mass loss of 31.7% is observed in the range of 222 °C to 428 °C due to the breakdown of Zn MOF, followed by a mass loss of 13.6% up to 800 °C, attributed to the formation of ZnO. Therefore, the thermal stability of Zn MOF is higher than that of Cu MOF. When evaluating the inclusion of Eu(III) in the system, it is found that Eu–Cu MOF exhibits higher thermal stability than Cu MOF, with a mass loss of 25.4% up to 505 °C due to the breakdown of the host framework, in addition to the release of absorbed solvent molecules. Similarly, Eu–Zn MOF is thermally more stable than Zn MOF, as evidenced by the higher mass loss of 27.4% observed up to 430 °C due to the breakdown of the structure.
In addition, PXRD analysis of Cu MOF and Zn MOF was performed to ensure the purity of the bulk sample, which coincides well with the simulated pattern obtained from single-crystal X-ray diffraction data (ESI,† Fig. S9 and S10). The PXRD pattern of Cu MOF exhibits diffraction peaks at 9.8°, 12.8°, 19.8°, and 25.5° corresponding to the (220), (400), (440), and (040) planes.37 From the comparison between the X-ray diffraction (PXRD) patterns of Cu MOF and the Eu–Cu MOF, additional diffraction peaks in the latter are identified at 10.6°, 18.5°, and 20.2°, along with the characteristic peaks of Cu MOF. These peaks can be tentatively assigned to the (010), (020), and (110) diffraction planes related to Eu(III) inserted into the structure (ESI,† Fig. S11). The PXRD of the Eu–Zn MOF also shows similar peaks at 9.2° and 19.1°, which are tentatively assigned to the (010), (020) diffraction planes, and other peaks at 2θ values of 11.2°, 14.2°, 17.9°, 25.1°, 31.2°, 42.3°, corresponding to the (002), (010), (012), (211), (006), and (008) crystal planes, confirming the presence of ligand-coordinated Eu(III) in the system (ESI,† Fig. S12). These assignments corroborate the reported literature on Eu(III)-infused MOFs with similar environments but different ligand backbones.38,39
Surface morphologies of Cu MOF investigated by field-emission scanning electron microscopy (FE-SEM) reveal a flat sheet or film-like morphology of the material; however, deposition of flakes is observed on the flat surface in Eu–Cu MOF after infusion (Fig. 4a–d). Zn MOF also reveals a similar sheet-like morphology and dot-like deposits on the sheets in Eu–Zn-MOF (Fig. 4e–h). Further EDAX analysis and elemental mapping confirmed the presence of Cu, C, O, N and Cu, C, O, N, Eu, Cl as constituent elements in Cu MOF and Eu–Cu MOF, respectively (ESI,† Fig. S13 and S14). Similarly, Zn, C, O, N, and Eu, Cl, Zn, C, O, N were observed as constituent elements in Zn MOF and Eu–Zn MOF, respectively (ESI,† Fig. S13 and S15).
![]() | ||
Fig. 4 FE-SEM images of (a and b) Cu MOF, (c and d) Eu–Cu MOF, (e and f) Zn MOF, and (g and h) Eu–Zn MOF. |
The XPS survey spectra of all the materials support the results of EDAX and the presence of respective elements in Cu MOF, Zn MOF, Eu–Cu MOF, and Eu–Zn MOF (Fig. 5). In Cu MOF, the spin–orbit pairs Cu 2p1/2 and Cu 2p2/3 with binding energy peaks at 953.1 eV and 933 eV correspond to the +2 oxidation state with additional satellite peaks at 943.5 eV and 963 eV (ESI,† Fig. S16a). Also, the XPS spectra of C 1s, O 1s, and N 1s of Cu MOF are as expected (ESI,† Fig. S16b–d). The XPS spectra of Eu–Cu MOF show the photoelectron peaks at 1164.1 eV and 1134.6 eV due to the contribution of spin–orbit pairs 3d3/2 and 3d5/2, which indicates the successful encapsulation of Eu(III) in the host framework (ESI,† Fig. S17a). In addition, the binding energy peaks observed at 953.4 eV and 933.1 eV correspond to the +2 oxidation state of Cu with additional satellite peaks at 943.3 eV and 963.1 eV (ESI,† Fig. S17b). Eu–Cu MOF also retained the characteristic peaks of Cl 2p, O 1s, N 1s, and C 1s in the Eu–Zn MOF framework (ESI,† Fig. S17c–f). In Zn MOF, the Zn 2p scan shows two major characteristic peaks with binding energies of 1043.1 eV and 1020 eV, attributed to the spin–orbit pairs of 2p1/2 and 2p3/2 of Zn2+ ions, respectively (ESI,† Fig. S18a). Additionally, the XPS spectra of O 1s, N 1s, and C 1s of Zn MOF are found as expected (ESI,† Fig. S18b–d). The XPS spectra of Eu–Zn MOF show the photoelectron peaks at 1163.3 eV and 1133 eV due to the contribution of spin–orbit pairs 3d3/2 and 3d5/2, which indicates the successful encapsulation of Eu(III) in the Eu–Zn MOF framework (ESI,† Fig. S19a). The binding energy peaks at 1043.6 eV and 1020.5 eV, corresponding to the presence of Zn2+, were retained along with the characteristic peaks of Cl 2p, O 1s, N 1s, and C 1s (ESI,† Fig. S19b–f).
The photoluminescence spectra of SB and MOFs were recorded at room temperature in methanol, and the spectra are given in Fig. 6. SB exhibits an emission band centred at 491 nm, and another at higher energy at 457 nm when excited at 400 nm, related to the fluorescence and phosphorescence of this ligand.41,42 The Cu MOF spectrum displays an emission band centred at 451 nm and another with a maximum at 426 nm, with higher energy, also under excitation at 400 nm, resulting in an emission colour in the blue region, as verified by its CIE coordinate (0.17, 0.12) and CCT magnitude of 3931 K (Fig. 7). Notably, the luminescence of Eu(III) complexes can be sensitized upon the near-visible light excitation (350 nm < λex < 415 nm) and 465 nm. The emission spectra of Eu–Cu MOF and Eu–Zn MOF were recorded in this range of wavelengths; however, excitations at λ < 360 nm and λ > 410 nm failed to produce white light emission and gave the perception of blue, turquoise blue, and green for Eu–Cu MOF, and blue, green, and yellow for Eu–Zn MOF, respectively (ESI,† Fig. S21a, b and S22). Therefore, the PL spectra were recorded in the region 370–400 nm, and the best results were obtained at 400 nm.43 In contrast, in the emission spectrum of Eu–Cu MOF, compared to that of the precursor, the bands in the blue region underwent a bathochromic shift, being now observed at 483 nm and 456 nm, together with the characteristic Eu(III) bands centred at 591 nm (5D0 → 323 7F1), 617 nm (5D0 → 7F2), 694 nm (5D0 → 7F4). In this case, the visual perception of the emission appears green, with a CIE 1931 coordinate of (0.25, 0.35), CCT magnitude of 9896 K, and CRI value 76 (Fig. 6 and 7). The photoluminescence quantum yield (PLQY) of Eu–Cu MOF calculated using standard rhodamine 6G is found to be 74.4%.44 However, for the Eu–Cu MOF, the combinations of the emission bands, considering the variation of their relative intensities, did not result in the perception of white light for certain excitation wavelengths.
Zn MOF at the same excitation wavelength of 400 nm shows an emission band centred at 464 nm displaying blue light emission with the CIE coordinate (0.18, 0.22) and the corresponding CCT magnitude of 145506 K (Fig. 7a). In contrast, Eu–Zn MOF displays a blue colour upon excitation at 370 nm, a turquoise blue colour at 380 and 390 nm, and a near-white light emission at 400 nm (Fig. 7c). The most effective tunability of emission bands was observed at an excitation wavelength of 400 nm, where the intensity of emission bands was the maximum resulting in near white light emission. The spectrum displays emission bands centred at 481 nm, along with characteristic bands due to Eu(III) ions centred at 697 nm (5D0 → 7F4), 615 nm (5D0 → 7F2), 590 nm (5D0 → 7F1) displaying near white light emission with CIE coordinates (0.31, 0.29), a CCT magnitude of 6719 K, and CRI value of 73 and PLQY equal to 84.9% (ref. 44) (Fig. 7a). The CIE coordinates lie close to the NTSC (National Television Standards Committee) value. However, the CRI values are lower than the ideal threshold (>80) typically desired for high-quality white light applications.
The electronic transition mechanism involved in Eu–Cu MOF and Eu–Zn MOF is due to the narrowed band characteristics of Eu(III) ions observed from its excited level 5D0, which may also be populated by intramolecular energy transfer (IET) via the triplet state of the ligand, T1 {T1 → Eu(III)}, which in turn is the result of the intersystem crossing (ISC) from the singlet state S1 to T1 {S1 → T1}, also known as the “antenna effect” (ESI,† Fig. S23).45 The encapsulated MOFs with donor-type ligands strongly interact with the 4f orbitals of the dopant and provide a rigid structural framework, thereby enhancing charge-transfer (CT) emissions from π-orbitals to 4f orbitals.
Although these materials do not meet this benchmark, the observed CRI values are comparable to those of early-stage phosphor-based systems, indicating the potential for improvement through further material engineering. The observed CRI values may be attributed to an incomplete spectral balance despite europium doping. It is hypothesized that enhancement could potentially be achieved through ligand engineering to broaden the emission profile or by introducing co-ligands with complementary emissive properties. Such modifications may improve spectral coverage and consequently, the CRI, without compromising luminescent performance. As compared to previous reports on Er–Yb-MOF and Er–Eu-MOF by Xu et al. and on a high-efficiency perovskite emitter (Cs3Cu2X5) (X = Cl, Br, and I) by Chen et al., the current report offers a new class of structurally tunable, ligand-directed white light-emitting materials that operate under low voltages and emit light across a wide spectral range. Although the PLQY of the hybrids is lower than that of Cs3Cu2X5 and does not exhibit upconversion like Ln-MOFs, they present a balance of optical tunability, chemical versatility, and ease of synthesis, with potential for further development in single-component white-light applications.46,47
![]() | ||
Fig. 8 Display of (a) non-MOF-LED off state, (b) Eu–Cu MOF-LED off state, (c) Eu–Zn MOF-LED off state, (d) non-MOF-LED on state, (e) Eu–Cu MOF-LED on state, and (f) Eu–Zn MOF-LED on state. |
To evaluate the emission retention of Eu–Zn MOF, emission spectra of the coated material were recorded after a period of 1, 3, 5, and 7 days. The result shows that all emission bands retain their identity throughout the period, indicating good emission retention (ESI,† Fig. S24). To evaluate the humidity resistance and photostability of Eu–Zn MOF, a coating of the MOF was applied onto the plastic epoxy resin cover of a commercially available LED. The coated LED was exposed to ambient air and continuously operated (switched on) for durations of 1, 2, and 3 days. After each exposure period, the outer MOF coating was analyzed using FTIR spectroscopy to monitor any changes. The FT-IR analysis reveals that the structural framework and chemical composition of the composite remain largely unchanged, thus depicting stability under prolonged light exposure and humid conditions (ESI,† Fig. S25). This method of tuning the properties of MOF with luminescence active centres {Eu(III)} also provides a new direction for the fabrication of high-performance and cost-effective light-emitting devices.
Footnote |
† Electronic supplementary information (ESI) available: Characterization data of all the compounds, elemental mapping, XPS spectra, PXRD pattern of Cu MOF, Zn MOF, Eu-Cu MOF, Eu-Zn MOF. Crystallographic data and structural parameters; crystallographic data are comprised in the CCDC codes of 2444353 (Zn MOF) and 2444361 (Cu MOF). For ESI and crystallographic data in CIF or other electronic format see DOI: https://doi.org/10.1039/d5ce00448a |
This journal is © The Royal Society of Chemistry 2025 |