Diksha
Malik
ab and
Srinivasan
Natarajan
*a
aFrameworks Solid Laboratory, Solid State and Structural Unit, Indian Institute of Science, Bangalore – 560012, India. E-mail: rajan1960@gmail.com
bInorganic and Physical Chemistry department, Indian Institute of Science, Bangalore – 560012, India
First published on 7th October 2025
New transition metal substituted milarite silicates of the general formula A2B2C[T(2)3T(1)12O30] were prepared by a conventional solid state technique and their structures determined by powder X-ray diffraction (PXRD) methods. Raman spectroscopic studies indicated expected Raman bands. The oxidation states of the transition elements were confirmed by XPS studies. The optical absorption bands were rationalized based on the ligand-centered emission and allowed d–d transitions. The white compounds exhibited good deep UV cut off values, suggesting their possible use as optical filters. The white compounds also showed good near-infrared (NIR) reflectivity comparable to that of the commercial TiO2. The rare earth substituted compounds, Na2Mg2.5Ca0.5−xLnxZn2Si12O30, exhibited intense red (Eu3+), green (Tb3+) and blue (Tm3+) emissions. An optimal composition having all the three lanthanide ion substitutions resulted in white light emissions in Na2Mg2.5Ca0.48Tm0.01Tb0.02Eu0.01Zn2Si12O30. The presence of Co2+ in the Na2Mg3Co2Si12O30 compound facilitated the study of the oxygen evolution reaction (OER) with good values that are comparable to those of RuO2. The white compounds have reasonably good dielectric constant values with low dielectric loss, which indicates their possible use in communication devices. The milarite based compounds exhibited considerable potential towards new colored compounds, white light emission and OER properties. The present study suggests that it is profitable to explore mineral structures towards new and known material properties.
Another important class of compounds belong to the silicates. The ease of formation of polymeric structures from [SiO4] tetrahedra results in a large variety of interesting structures.10–12 Many silicates occur naturally and a number of synthetic ones are also known, especially those of zeolites.13–15 The structural variety and diversity of the silicates, notably the porous ones, find many applications in the area of catalysis, gas sorption and separation processes.
Some of the naturally occurring silicates have alkali/alkaline earth metal ions as a part of the structure.16–19 One such naturally occurring silicate belongs to the milarite family of compounds. The general formula of the milarite structure can be written as A2B2C[T(2)3T(1)12O30]·xH2O,20,21 where T refers to atoms having tetrahedral coordination, A refers to atoms having octahedral coordination, B refers to atoms having 9-coordination and C refers to atoms having 12-coordination. This diversity in accommodating different coordinations within the structure results in a large variety of structural modifications within the milarite family of compounds.20,22–27 The variety of ions that can be accommodated within the milarite structure is given in Table S1.28
The structure of the milarite consists of T(1) [SiO4] tetrahedra that form a six-membered ring, which are connected to form a double six ring of [Si12O30] units. The double six rings are joined by the T(2) tetrahedra forming the extended structure (Fig. S1). The T(2) tetrahedra share edges with A-octahedra, which are also connected with T(1) tetrahedra, strengthening the overall tetrahedral framework. The B-polyhedra occupy positions above and below the A-octahedra. Both the A and B sites occupy a 3-fold axis and exhibit considerable distortions. The C-atoms are in the channels formed by the connections between the A, B, T(1) and T(2) polyhedral units.
We have been interested in the study of many mineral structures recently with a view to explore their physical and chemical properties.29–36 The milarite structure is appealing to us due to the possibility of exploring different substitutional studies at various positions. Thus, we have explored the milarite structure towards the generation of new colored compounds as a host for luminescence and white light emission and towards oxygen evolution reactions. In this paper, we present the results of the studies on the compounds prepared with the milarite structure.
![]() | ||
| Fig. 1 Experimental PXRD patterns of the prepared compounds, Na2Mg5Si12O30, Na2Mg3Zn2Si12O30, Na2Mg3Co2Si12O30, and Na2Mg2.5Ca0.5Zn2Si12O30. | ||
The PXRD patterns were recorded employing a Philips X'pert diffractometer (Ni filtered Cu Kα radiation, λ = 1.5418 Å) (Fig. 1 and Fig. S2). For Rietveld refinement of Na2Mg3Co2Si12O30 and Na2Mg3ZnCuSi12O30, the PXRD data were collected at room temperature in the 2Θ range of 8–120° with a step size of 0.02° and a step duration of 50 s. The PXRD pattern was refined using GSAS-EXPGUI. The lattice parameters, scale factor, background (Fourier polynomial background function), pseudo-Voigt (U, V, W, and X), and isothermal parameter (Uiso) were refined (Fig. 2, Table 1, Fig. S3 and Table S2). Thermal parameters were constrained to be the same for different atoms occupying the same site. The optical studies were carried out on the powdered samples (Perkin–Elmer Lambda 950 UV-Vis double-beam spectrometer) in the spectral region of 200–2500 nm. The diffuse reflectance data were converted to the Kubelka–Munk unit:
| F(R) = (1 − R)2/2R = α/S |
| αhν = A[hν − Eg]n |
![]() | ||
| Fig. 2 Rietveld refinement of Na2Mg3Co2Si12O30. The observed (O), the calculated (red line), and the difference (bottom blue line) are shown. The vertical bars (|) indicate Bragg's reflections. The simulated pattern was generated from Na2Mg5Si12O30.26 | ||
| Elements | x/a | y/b | z/c | Occ. | U iso | Site |
|---|---|---|---|---|---|---|
| a Space group P6/mcc; a = b = 10.1537(1) Å; c = 14.2536(1) Å; V = 1272.65(5) Å3; α = β = 90°; γ = 120°; reliability factors: RP = 1.2%; Rwp = 1.9%; χ2 = 4.80; average bond lengths [Å]: Mg2/Co2–O = 2.0410(1); Si–O = 1.5854(1); Mg1–O = 2.1324(1). b The refinement structural model was taken from Na2Mg5Si12O30.26 | ||||||
| Na1 | 0 | 0 | 0.2500 | 0.670 | 0.084(10) | 2a |
| Na2 | 0.3333 | 0.6666 | 0 | 0.670 | 0.023(4) | 4d |
| Mg1 | 0.3333 | 0.6666 | 0.2500 | 1 | 0.012(3) | 4c |
| Mg2 | 0.5000 | 0 | 0.2500 | 0.363 | 0.037(2) | 6f |
| Co2 | 0.5000 | 0 | 0.2500 | 0.637 | 0.037(2) | 6f |
| Si1 | 0.7641(3) | 0.1117(3) | 0.1077(2) | 1 | 0.018(1) | 24m |
| O1 | 0.7385(6) | 0.1370(7) | 0 | 1 | 0.024(3) | 12l |
| O2 | 0.9380(6) | 0.2145(5) | 0.1301(3) | 1 | 0.019(2) | 24m |
| O3 | 0.6674(5) | 0.1593(4) | 0.1667(2) | 1 | 0.017(2) | 24m |
The dielectric measurements (Novocontrol impedance analyzer (Alpha-A)) were carried out in the frequency range of 0.1 Hz–1 MHz. Micro-Raman spectral studies (HORIBA LabRAM HR Evolution) were performed at room temperature using a triple monochromator DXR microscope Raman spectrometer, equipped with 1800 groove per mm diffraction grating and a liquid nitrogen-cooled CCD detector. The 10 mW output of the 532 nm line of Nd:YAG (10 mW) was used as the excitation source.
Magnetic measurements were carried out in the temperature range 2–300 K employing a SQUID magnetometer (Quantum Design, USA) under an applied field of 1000 Oe. The photoluminescence and lifetime studies (Fluoromax plus spectrofluorometer and Edinburgh spectrofluorometer) were performed at room temperature.
The electrocatalytic performance of Na2Mg3Co2Si12O30 (PARSTAT analytical electrochemical workstation) was studied employing a three-electrode system with Hg/HgO (0.5 M KOH) as the reference electrode, a graphite rod as the counter electrode, and glassy carbon coated with Na2Mg3Co2Si12O30 ink as the working electrode in a 0.5 M KOH electrolyte.
We made attempts to refine the Co2+ ion substituted compound, Na2Mg3Co2Si12O30, by employing the Na2Mg5Si12O30 compound as the model using the GSAS-EXPGUI powder structure refinement program (Fig. 2). The results of the refinement indicate a good fit with the reported structure, Na2Mg5Si12O30.26 A minor peak observed around ∼22° appears to be due to a SiO2 phase, which has been previously reported in the synthesis of milarites.26 This small peak was excluded during the refinement. The refinement suggested that the Co2+ ions are substituted at the tetrahedral sites (T(2)) and resulted in a small reduction in the overall volume of the unit cell (1275.03(5) Å to 1272.65(5) Å). This reduction in the overall volume appears to be consistent with the contraction of the c-axis (14.29(3) Å to 14.25(5) Å), reflecting the stronger polarizing power exhibited by the Co2+ ions compared to the Mg2+ ions in the tetrahedral coordination (Fig. 2, 3 and Table 1). We carried out the bond valence sum (BVS) calculation on the compounds. We observed that the BVS value for the Co2+ ions at the tetrahedral T(2) site was ∼1.56, which suggests that the tetrahedral site may be slightly under-bonded. In the present compound, Na2Mg3Co2Si12O30, the BVS calculations reveal that the octahedral A site (Mg2+ ion) gives a value of ∼1.85 and the T(2) site (Mg2+ ions) was found to have a value of ∼1.548. This indicates that the T(2) tetrahedral positions in the milarite structure are inherently under-bonded. The BVS calculations for Co2+ ions showed that they were under-bonded, and it is likely that the Co2+ ion occupies the T(2) site rather than the A site.
Of these, 3 vibrational modes (A2u + E1u) are acoustic in nature, and 67 modes (13A2u + 27E1u) are IR-active. Based on the nuclear site analysis, the following Raman active modes were to be expected.
| ΓRaman = 11A1g + 26E2g + 24E1g |
We observed a lower number due to the doubly degenerate Raman modes and overlap of bands. A weak band at ∼567 cm−1 could be due to the bending vibration of Si–O bonds within the SiO4 tetrahedra.40 A weak broad band observed at ∼823 cm−1 might be due to the Si–O–Si vibrations, and is commonly observed in many silicates.26 The bands at ∼1050 and ∼1121 cm−1 (A1g) could be due to the stretching vibration of the Si12O30 cluster.24,25
The Mg(1)O6 bending vibration was observed at ∼241 cm−1.41 The substitution of Co2+ ions in place of Mg2+ ions results in a Raman splitting of the band centred around 240–275 cm−1 and transforms into a doublet. In addition, we also note that the Raman band centred at ∼176 cm−1 appeared to gain intensity due to the substitution of Co2+ ions. This suggests a stronger bonding and more polarizability in the Co-substituted compound. The medium Raman bands centred at ∼511 cm−1 (E2g) and ∼567 cm−1 could be due to the Mg(2)–O4 stretching vibrations. The substitution of Co2+ ions in place of the tetrahedral Mg2+ ions (Mg(2)) leads to the following changes in the Raman spectra: (i) the band centred at ∼511 cm−1 increases in intensity as the amount of Co2+ ions is increased and (ii) the band centred at ∼567 cm−1 shifts to a lower wavenumber with the increase in the substitution of Co2+ ions. The strong Raman band observed at ∼467 cm−1 could be due to the stretching vibration of Mg–O–Si bonds (Si12O30 units binding with Mg), which appears to be a characteristic marker of the milarite family of compounds.26
The Co2+ ions (3d7) possess 4F and 4P states as the free ion state. On application of the octahedral or tetrahedral ligand field, these terms split into multiple levels, which results in the characteristic d–d transitions.45 Normally, the Co2+ ions in a tetrahedral coordination environment have three spin-allowed transitions: 4A2 → 4T2 (4F), 4A2 → 4T1 (4F), and 4A2 → 4T1 (4P) (Fig. S6a). Out of these, the first two transitions are observed in the infrared (IR) region and the third transition, 4A2 → 4T1 (4P), is observed in the visible region. In many compounds, this allowed transition is often found overlapped with the spin-forbidden transitions such as 4A2(4F) → 2E(2G).46–48 In the UV-Vis spectra, the compound, Na2Mg5−xCoxSi12O30 (x = 0.25–2), exhibits two maximas that can be separated based on the Co2+ ion concentration in the compounds. Thus at the low substitution region (x = 0.25–0.5) the maxima is observed at ∼2.03 eV (∼610 nm) and ∼2.48 eV (∼500 nm) with a shoulder at 2.13 eV (∼580 nm). These transitions correspond to the 4A2 → 4T1 (4P) transition of the Co2+ ions in tetrahedral coordination.49 In addition we also observed the spin forbidden transitions at 4A2(4F) → 2E(2G) (a doublet at ∼2.98 eV (∼416 nm) and ∼3.02 eV (∼410 nm)) and 4A2(4F) → 2T1(2G) (∼2.31 eV (∼536 nm)). Similar earlier observations were rationalised based on the highly distorted Co2+ ions in tetrahedral coordination along with strong L–S coupling.50 The absorptions for the Co2+ ions were observed in the red and green-blue regions, which would result in a lavender color. When the concentration of the Co2+ ions increases (x = 1–2), the strong absorption band observed at ∼2.03 eV reduces in intensity along with an increase and broadening of the absorption at ∼2.48 eV. This results in the deepening of the lavender color.51 In addition, two valleys of low absorption around ∼2.3 eV and ∼2.88 eV, corresponding to the yellow-green and deep blue regions, were also observed. A combination of the absorption bands along with valleys of low absorption would result in the observed color of the compounds containing higher concentration of Co2+ ions.51 The observation of a triplet broad band in the region between 0.88 and 1.02 eV, which corresponds to the 4A2 → 4T1 (4F) transition, suggests the presence of Co2+ ions in a tetrahedral coordination environment (Fig. 5).52,53 We also observed similar absorption bands for the Co2+ ion substituted Na2Mg3ZnCoSi12O30 compound.
The Cu2+ ion substituted Na2Mg4−xCuxSi12O30 (x: 0.5, 1) and Na2Mg3Z2−xCuxSi12O30 (x: 0.5, 1) compounds exhibit colors that have different shades of blue under the daylight (Fig. S6). The UV-visible spectra were featureless. We observed a broad transition in the near infrared (IR) region (1.34 eV–1.51 eV) which could be due to the 2T2 → 2E2 transition. Similar optical absorption features have been observed for the Cu2+ ions in oxide hosts.33,36
The Ni2+ ion substituted compounds, Na2Mg4.5Ni0.5Si12O30 and Na2Mg3Zn1.5Ni0.5Si12O30, exhibited a greyish green color under the daylight (Fig. S6). We observed a broad absorption band with a maximum centred at ∼1.96 eV (∼632 nm) and ∼2.44 eV (∼508 nm) in the UV-Visible spectrum. Normally, Ni2+ ions prefer either a square-planar or an octahedral coordination. The Ni2+ ions substituted in tetrahedral coordination are also known and the colors exhibited by these compounds change from pink to grey color.31–33,35,51,54 In the present compound, we successfully substituted a small concentration of Ni2+ ions in the tetrahedral coordination. The Ni2+ (d8) ions in a tetrahedral environment are expected to exhibit three transitions, 3T1 (F) → 3T2 (F), 3T1 (F) → 3A2 (F), and 3T1 (F) → 3T1 (P). Of these, the 3T1 (F) → 3T1 (P) transition corresponds to the observed absorption band.33,35,36
As part of the study, we also carried out the NIR reflectance studies on the white-colored compounds (Fig. S7). All the white compounds exhibited good near-IR reflectivity behavior with values in the range of 70–85%, which is comparable to that of the industrial standard NIR compound, TiO2 (∼70%) (Fig. S7). The high near-IR reflectivity values suggest that these compounds may find use as NIR reflective coatings.35
There has been intensive research towards developing materials that exhibit deep ultraviolet (UV) cut-off characteristics, which are useful for advanced optical applications such as in UV filters, protective coatings, optical components in lithography and sterilization technologies etc. It has been shown that many borates, phosphates, and fluorides exhibit a short UV cutoff edge.55–57 We explored the compounds, Na2Mg5Si12O30 and Na2Mg3Zn2Si12O30, towards the deep UV-cutoff (Fig. S8). The investigations indicate that the compounds exhibit ∼72% of the UV cut-off at 204 nm. This value appears to be promising towards the development of optical components requiring a deep UV-cut off.
All the lanthanide substituted compounds were found to be single phasic by PXRD (Fig. S2). The room temperature photoluminescence of all the compounds exhibited the expected emission consistent with the lanthanide ions (Fig. 6 and Fig. S9). In all the compounds, the lanthanide concentrations were varied to establish maximum emission intensity. The studies clearly establish that for Eu3+ ions, the maximum intensity was observed at 6%; for Tb3+ ions it was at 9% and for Tm3+ it was at 4% (Fig. S9).
![]() | ||
| Fig. 6 PL and PLE spectra of Na2Mg2.5Ca0.5−xLnxZn2Si12O30: (a) Eu3+ (6%), (b) Tb3+ (9%), and (c) Tm3+ (4%). | ||
The photoluminescence excitation (PLE) spectra of the 6% Eu3+ ion substituted compound, Na2Mg2.5Ca0.44Eu0.06Zn2Si12O30, were monitored at 612 nm (Fig. 6a). This exhibited excitation peaks at 361 nm, 381 nm, 394 nm, 412 nm, and 464 nm, which correspond to the f–f transitions, originating from the ground state 7F0 to 5D4, 5L7, 5L6, 5D3, and 5D2 excited states, respectively.30,61 Additionally, we also observed an excitation peak at 531 nm, which may correspond to the 7F1 → 5D1 transition.29 The emission spectra (λex = 394 nm) indicated a maximum intensity at 612 nm wavelength, which could be due to the 5D0 → 7F2 transition. The other emission peaks were observed at 588 nm, 649 nm and 702 nm in the photoluminescence emission (PL) spectra, which correspond to the 5D0 → 7F1, 5D0 → 7F3, and 5D0 → 7F4 transitions, respectively. It is known that the emission of Eu3+ ions in a host can provide clues to the local environment, as the emission characteristics are sensitive to the site symmetry and the coordination. In the present compounds the maximum intensity was observed at 612 nm, assigned to the 5D0 → 7F2 transition, which is the electric dipole transition. This transition exhibits enhanced intensity when the Eu3+ ions occupy non-centrosymmetric or low-symmetry sites. The magnetic dipole transition, 5D0 → 7F1, on the other hand, is parity-allowed and relatively insensitive to the changes in the crystal field. The observation of maximum intensity for the 5D0 → 7F2 transition (612 nm) suggests that the Eu3+ ions are located at sites lacking the inversion symmetry and corroborates the low-symmetry environment.61
The PLE and PL spectra of the 9% Tb3+ ion substituted compound, Na2Mg2.5Ca0.41Tb0.09Zn2Si12O30, were monitored at 541 nm and 377 nm, respectively (Fig. 6b). The PLE spectra of the Tb3+ ion substituted compound exhibited excitation peaks at 319 nm, 333 nm, 351 nm, 367 nm, 377 nm, and 483 nm, which correspond to the excitation from the ground state, 7F6, to the excited states: 6D1, 5L7,8, 5L10, 5G6, 5D3 and 5D4, respectively.62,63 The PL spectrum of the compound (λex = 377 nm) exhibits emission bands in the blue region at 436 and 486 nm, the green region at 541 and 548 nm, and the orange-red region at 583 and 619 nm.64,65 Of these, the most intense emission was found to be at 541 nm, which could be due to the 5D4 → 7F5 transition.66 The other emission bands observed in the PL spectra could be assigned to 5D4 → 7F6 (486 nm), 5D4 → 7F4 (583 nm), and 5D4 → 7F3 (619 nm) transitions. The weak emission observed at ∼436 nm could be due to the 5D3 → 7F4 transition.64–67
The PL and PLE spectra of the compound, Na2Mg2.5Ca0.46Tm0.04Zn2Si12O30, doped with 4% Tm3+ ions were recorded by monitoring emissions at 357 nm and 457 nm, respectively (Fig. 6c). It is known that the Tm3+ ions exhibit many transitions due to the strong Russell–Saunders (L–S) coupling of the 4f electrons.68 The PLE spectra exhibited a strong excitation peak at 357 nm, which could be due to the 3H6 → 1D2 transition. The Tm3+ substituted compound exhibits an intense blue emission at 457 nm (λex = 357 nm), which can be assigned to the 1D2 → 3F4 transition.58,61,68,69
The Commission Internationale de L'Eclairage (CIE) chromaticity coordinates for the substituted compounds exhibit intense blue (4% Tm3+), green (9% Tb3+) and red (6% of Eu3+) emission (Fig. S10). The corresponding color correlating temperature (CCT) values are given in Table S4.
The observation of RGB emission in the Na2Mg2.5Ca0.5Zn2Si12O30 compound suggested the possibility of obtaining white light by careful choice of the substitution of the three rare-earth ions (Table S5a). The CIE coordinates and the CCT values were also found to be comparable to those of some of the known silicate compounds (Table S6). It may be noted that the ratio of the Tm3+
:
Tb3+
:
Eu3+ ions in the host exhibiting maximum intensity was found to be approximately 1
:
2
:
1. To this end, we synthesized a series of compounds by fixing the concentrations of Tm3+ (1%) and Tb3+ (2%), along with systematically varying the Eu3+ ion concentration (x = 0, 0.5, 1, 1.5, 2 mol%) (Fig. 7). The photoluminescence (PL) spectra of the different samples were recorded by employing an excitation wavelength of 358 nm (Fig. 7). The variation of Eu3+ ion concentration in the host Na2Mg2.5Ca0.5Zn2Si12O30 (1% Tm3+, 2% Tb3+, and x% Eu3+ with x = 0–2%) changes the color of the emission from the blue green region to the white region (Fig. 8). In addition, we also noted that on increasing the Eu3+ concentration from 0 to 2 mol%, there was a progressive quenching of the emission intensities of the Tm3+ and the Tb3+ ions, accompanied by an enhancement of the intensity of the Eu3+ emission. This indicates possible energy transfer from the Tm3+ and Tb3+ ions to the Eu3+ ions. Similar behavior has been observed before.30,61 The PL Tm3+ emission at ∼457 nm (blue) and the Tb3+ emission at ∼541 nm (green) are progressively quenched as the concentration of Eu3+ substitution increases (Fig. 7). The change in the intensities of the individual emission indicates that there is a possibility of cascade energy transfer (Tm3+ → Tb3+ → Eu3+).
![]() | ||
| Fig. 8 CIE diagram for Na2Mg2.5Ca0.5Zn2Si12O30 (1% Tm3+, 2% Tb3+, and x% Eu3+) with the concentration of Eu3+ varying from x = 0 to x = 4%. Note the white light region for x = 1. | ||
We have carried out a temperature-dependent photoluminescence on a freshly prepared and characterized white light emitting sample (Fig. S9d). The CIE coordinates obtained at 20 K indicate that the values are marginally different from those obtained at room temperature (Table S5b).
![]() | ||
| Fig. 9 PL emission spectra (λex = 377 nm) of Na2Mg2.5Ca0.5Zn2Si12O30 (8% Tb3+ and x% Eu3+), x = 2 to x = 6%. | ||
Thus, the critical distance (Rc) for the Tb–Eu pair in the present structure was calculated to be ∼21.6 Å. This value is considerably larger than the 5 Å required for the exchange interactions and suggests that the energy transfer occurs through a long-range interaction (multi-ploar interaction). It is likely that the energy transfer between the Tb3+ ions and the Eu3+ ions could be through the resonant electric dipole–dipole coupling pathway. This is also consistent with the observation that the Tb3+ ion emission intensity is substantially reduced in the presence of a higher concentration of Eu3+ ions. It may be noted that the Eu3+ (5D0 → 7F2) emission intensity is increased which indicates that the Tb3+ ions act as a good sensitizer for the Eu3+ ions through long-range dipole–dipole interactions resulting in the energy transfer (Fig. 10).74,75,79–81
![]() | ||
| Fig. 10 Possible energy transfer pathway in the compound, Na2Mg2.5Ca0.5Zn2Si12O30 (1% Tm3+, 2% Tb3+, and x% Eu3+) with x = 0 to x = 2%. | ||
| τavg = (B1τ12 + B2τ22 + B3τ32)/(B1τ1 + B2τ2 + B3τ3) |
![]() | ||
| Fig. 11 Fluorescence lifetime decay curves (λex = 377 nm; λem = 541 nm) of Na2Mg2.5Ca0.5Zn2Si12O30 (8% Tb3+, and x% Eu3+) with the concentration of Eu3+ varying from x = 2 to x = 6%. | ||
The fluorescence average lifetime decay was also estimated for a single substituent having maximum emission intensity: 6% Eu3+, 9% Tb3+, and 4% Tm3+ in the compound Na2Mg2.5Ca0.5Zn2Si12O30 (Fig. S11).
Similarly, the fluorescence decay measurements were also conducted on Na2Mg2.5Ca0.5Zn2Si12O30 (1% Tm3+, 2% Tb3+, and x% Eu3+) with the concentration of Eu3+ varying from x = 0 to x = 2%. All the samples were excited at 358 nm, and the emission was monitored at 541 nm, corresponding to the 5D4 → 7F5 transition of Tb3+ ions (Fig. 12).
The fluorescence decay profile was fitted employing the tri-exponential function and the average lifetime was calculated (Table S7). The decreasing lifetime decay agrees with the energy transfer from Tm3+ to Tb3+ to Eu3+ ions, where the excited-state energy of Tb3+ is non-radiatively transferred to the Eu3+ ions. We obtained the white light emission at x = 1 for Eu3+ ions in the host Na2Mg2.5Ca0.5Zn2Si12O30 (1% Tm3+, 2% Tb3+, and x% Eu3+), which exhibited a fluorescence average lifetime decay of 485 μs. The fluorescence decay studies on the Na2Mg2.5Ca0.5Zn2Si12O30 (1% Tm3+, 2% Tb3+, and 1% Eu3+) compound at 20 K gave a lifetime value of 783 μs (Fig. S11d).
To evaluate the performance of the phosphor, we also measured photoluminescence quantum yield (PLQY) for the maximum La3+ substituted, Na2Mg2.5Ca0.5Zn2Si12O30, and for the white light compound, Na2Mg2.5Ca0.5Zn2Si12O30 (1% Tm3+, 2% Tb3+, and 1% Eu3+). The Eu3+ substituted compound, Na2Mg2.5Ca0.44Eu0.06Zn2Si12, exhibited a PLQY of 55% (λex = 394 nm). The Tb3+ substituted compound, Na2Mg2.5Ca0.41Tb0.09Zn2Si12, and Tm3+ substituted compound, Na2Mg2.5Ca0.46Tm0.04Zn2Si12, exhibited lower PLQY values of 16% (λex = 377 nm) and 1% (λex = 358 nm), respectively. The white light emitting phosphor, Na2Mg2.5Ca0.5Zn2Si12O30 (1% Tm3+, 2% Tb3+, and 1% Eu3+), yielded a PLQY value of 5%. Similar values have been reported in the literature.30
The molar magnetic susceptibility (χm) increases with decreasing temperature, exhibiting a sharp rise below 50 K. The room temperature molar magnetic susceptibility was found to be 0.23 emu mol−1, which increases to a value of 1.38 emu mol−1 at 4 K. The inverse susceptibility (1/χm) data were fitted in two temperature ranges: (i) in the range of 50–150 K and (ii) in the range of 150–300 K, employing the Curie–Weiss law.84
The fitting yielded a Curie constant (C) of 18.03 emu mol−1 K and a Curie–Weiss temperature (θ) of −140.9 K in the temperature range of 150–300 K. The fit in the temperature range of 50–150 K gave a Curie constant (C) value of 11.03 emu mol−1 K−1 and a Curie–Weiss temperature (θ) of −26.8 K. The negative value of θ at both the temperature ranges suggests the presence of antiferromagnetic interactions between the Co2+ ions in the compound.
The effective magnetic moment (μeff) values were also calculated from the Curie constant using the relationship,
. The μeff values obtained are 5.98μB in the range 150–300 K and 4.685μB in the range 50–150 K, which are higher than the spin-only magnetic moment values of 3.87μB for the high-spin Co2+ ions (S = 3/2). The observed magnetic moment value appears to be marginally larger (150 K–300 K) than the spin-only values of the tetrahedral Co2+ ions. These larger magnetic moment values observed may be due to some orbital contribution, which is known for Co2+ ions (Fig. S12c).84
:
1 ratio. 5% Nafion was also added to the catalyst. The entire mixture was taken in an ethanol
:
water (2
:
3) solution and sonicated for 60 minutes. This solution was dropcast onto the glassy carbon electrode in small increments to ensure a uniform and homogeneous coating of the electrode surface. The electrode was dried under an infrared lamp to remove any residual solvent.
The reaction was carried out in 0.5 M KOH solution. The entire solution was bubbled with Ar gas for 30 min at room temperature to remove any dissolved oxygen before starting the reaction. The electrocatalytic reaction was carried out employing a three-electrode set up. The glassy carbon electrode, containing the catalyst, was the working electrode, mercury/mercuric oxide and graphite rod were employed as the reference and counter electrodes. The electrocatalytic studies were screened by linear scan voltammetry (LSV) with a scan rate of 10 mV s−1. The LSV indicated that the oxygen evolution was observed at a overpotential of 1.54 V vs. RHE. The LSV studies were also compared with those with RuO2, which exhibited a overpotential of 1.43 V vs. RHE (Fig. S14). The kinetics of the electrocatalytic behavior was obtained from the Tafel plot (Fig. 13).
![]() | ||
| Fig. 13 (a) Linear sweep voltammogram (LSV), (b) Tafel plot, (c) Nyquist plot, and (d) chronoamperometric studies on the Na2Mg3Co2Si12O30 compound. | ||
The slope for the Na2Mg3Co2Si12O30 compound was found to be 92.86 mV dec−1, which is marginally higher than that observed for RuO2 (90.37 mV dec−1) (Fig. 13b). These values suggest that the charge transfer kinetics of the Na2Mg3Co2Si12O30 compound is comparable to that of RuO2.
Electrochemical impedance spectroscopy (EIS) studies were carried out to understand the interfacial charge transfer behavior of Na2Mg3Co2Si12O30 during the oxygen evolution reaction (OER).93 The studies were carried out in the AC field with frequencies in the range of 0.1 Hz–1 MHz (Fig. 13c). The Nyquist plot exhibits typical behavior where the electrode processes involve both the charge transfer and double-layer capacitance effects.94,95 The experimental data were fitted employing an equivalent circuit model, the Randles’ circuit, comprising R1, R2, and CPE1. In this model, R1 corresponds to the uncompensated solution resistance, while R2 represents the charge transfer resistance at the electrode–electrolyte interface. The constant phase elements, CPE1, account for the non-ideal capacitive behavior (Table S8). The low value of R2 indicates reasonable charge transfer kinetics, which appears to be consistent with the electrocatalytic activity observed both in LSV and Tafel analysis. The reasonable agreement between the fitted curve and the experimental data suggests that the employed model is good (Fig. 13c and Fig. S15).
The long-term electrochemical stability was evaluated employing chronoamperometric measurements at a fixed applied potential (1.647 V vs. RHE). The compound exhibited good OER activity without a drop in current for up to 12 h (Fig. 13d). The LSV curves recorded before and after the 12-hour chronoamperometry test show minimal changes, indicating that the catalyst maintains its electrocatalytic activity over time. This consistency suggests that the material remains stable and effective for oxygen evolution reaction (OER) applications under prolonged operational conditions (Fig. 13d).
The XPS studies were carried out on the Na2Mg3Co2Si12O30 compound to learn about the oxidation state of Co2+ ions. The XPS spectra exhibited two peaks at ∼782.5 eV and ∼798.2 eV, which may be assigned to the 2p3/2 and 2p1/2 binding energies of Co2+ ions (Fig. S16a).29,54,96 As can be observed, in addition to the main peaks, a prominent satellite was also observed at ∼787.7 eV (∼5.2 eV higher than 2p3/2) for Co-2p3/2. This is suggestive of Co2+ ions in a high spin state.97–99 The separation between the 2p3/2 and 2p1/2 binding energies was observed to be ∼15.7 eV which also suggests the presence of Co2+ ions in the tetrahedral coordination.100
The XPS spectra of the Na2Mg3Co2Si12O30 compound was also recorded after 1000 cycles of cyclovoltammetry. We observed a similar behavior for the Co2+ ions with 2p3/2 and 2p1/2 peaks at ∼782.2 eV and ∼797.8 eV, respectively. As can be noted, the separation between the two XPS peaks was found to be ∼15.6 eV. This also indicates that there is no change in the coordination environment of the Co2+ ions in Na2Mg3Co2Si12O30. We observed a peak at ∼781 eV and at ∼796.4 eV, which is suggestive of the binding energies of 2p3/2 and 2p1/2 states of Co3+ ions. It is likely that the Co2+ ions undergo oxidation during the electrocatalytic reaction. The weak satellite peak observed at ∼791 eV is also suggestive of Co3+ (3d6). The ratio of Co2+/Co3+ can be calculated from the 2p3/2 peak, which indicates a value of ∼6.5
:
1 (Fig. S16b).
The electrochemically active surface area (ECSA) of Na2Mg3Co2Si12O30 was estimated by analyzing the capacitive current response in the non-faradaic region using cyclic voltammetry at varying scan rates (0.01–0.1 V s−1). The capacitive current density
vs. the scan rate (V s−1), gives a linear relationship (Fig. S17). The slope of the linear fit corresponds to the double-layer capacitance (Cdl), which was found to be 0.702 mF cm−2. Based on the standard specific capacitance value of 40 μF cm−2 for a flat surface in 0.5 M KOH, the ECSA was calculated using the relationship, ECSA = Cdl/Cs, resulting in a value of 17.6. The large surface area suggests that there are large numbers of electrochemically accessible sites at the interface.32,101
The turnover frequency (TOF) for Na2Mg3Co2Si12O30 was determined using linear scan voltammetry (LSV) performed at a scan rate of 10 mV s−1 in 0.5 M KOH (Fig. S18). The TOF quantifies the rate at which catalytic sites facilitate the conversion of the reactants to the products. Based on the measured current densities and the calculated number of active centers from the redox peak integration, the TOF at the relevant potential was determined to be 6.01 s−1 for a 10 mA cm−2 current density.32 Similar values have been reported before.102–104
The Cu2+ ions in the compound, Na2Mg3CuZnSi12O30, exhibited two main peaks at ∼933.3 eV and ∼952.7 eV, corresponding to the 2p3/2 and 2p1/2 binding energies of Cu2+ ions.105–107 Cu2+ ions exhibit strong satellite features in the XPS spectra, which are not generally observed for Cu+ ions.108 We observed strong satellite features at ∼9 eV higher for 2p3/2 and 2p1/2 at ∼942.2 eV and ∼961.6 eV, respectively. The difference between the two binding energies both for the main peaks and the satellite peaks was observed to be ∼19.3 eV, which is suggestive of Cu2+ ions in tetrahedral coordination.109
The XPS spectra of Ni-2p in Na2Mg4.5Ni0.5Si12O30 exhibited two main peaks at ∼856.3 eV and at ∼873.9 eV which correspond to the binding energies of 2p3/2 and 2p1/2 for Ni2+ ions, respectively. In addition, we also observed satellite peaks at ∼861.9 eV and ∼879.9 eV, which is also suggestive of the presence of Ni2+ ions. The separation between the two binding energies (2p3/2 and 2p1/2) was found to be ∼17.6 eV, which indicates the Ni2+ ions may be present in a tetrahedral coordination environment.54,110
![]() | ||
| Fig. 14 (a) The variation of the dielectric constant as a function of frequency and (b) the dielectric loss as a function of frequency. Note the small loss for all the compounds. | ||
| This journal is © The Royal Society of Chemistry 2025 |