Next Article in Journal
Current and New Approaches to Predict the Deflections of One-Way Flexural Members with a Focus on Composite Steel Deck Slabs Voided by Circular Tubes
Previous Article in Journal
Physicochemical Factors That Influence the Biocompatibility of Cationic Liposomes and Their Ability to Deliver DNA to the Nuclei of Ovarian Cancer SK-OV-3 Cells
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Time- and Temperature-Dependent Luminescence of Manganese Ions in Ceramic Magnesium Aluminum Spinels

1
N. S. Kurnakov Institute of General and Inorganic Chemistry, 31 Leninskiy Prospekt, 119991 Moscow, Russia
2
Istituto di Fisica Applicata “N. Carrara”, Consiglio Nazionale delle Ricerche, via Madonna del Piano 10, Sesto Fiorentino, 50019 Florence, Italy
3
Istituto Nazionale di Ottica, Consiglio Nazionale delle Ricerche, via Madonna del Piano 10, Sesto Fiorentino, 50019 Florence, Italy
4
P. N. Lebedev Physical Institute, 53 Leninskiy Prospekt, 119991 Moscow, Russia
*
Author to whom correspondence should be addressed.
Materials 2021, 14(2), 420; https://doi.org/10.3390/ma14020420
Submission received: 15 December 2020 / Revised: 11 January 2021 / Accepted: 12 January 2021 / Published: 16 January 2021
(This article belongs to the Special Issue Advances in Phosphors for Light-Emitting Diode)

Abstract

:
Samples of magnesium aluminum spinel ceramics doped with manganese ions were prepared by a high-temperature solid-state reaction method; their potential as red-emitting phosphors was analyzed using a time-resolved luminescence spectroscopy technique, from room temperature to 10 K. It was found that in the red spectral range, the luminescence spectra of manganese ions in the MgAl2O4 spinel showed a narrow band peaking at 651 nm due to the emission of Mn4+ and a broader emission band in the region of 675 ÷ 720 nm; the ratio of intensities for these bands depends on the synthesis conditions. By applying a special multi-step annealing procedure, the MgAl2O4:Mn4+ phosphor containing only tetravalent manganese ions, Mn4+, was synthesized. Broad-band far-red emission observed from MgAl2O4:Mn and Mg1.25Al1.75O3.75F0.25:Mn phosphors, prepared by a conventional method of a solid-state reaction, was interpreted as coming from Mn3+ ions.

1. Introduction

At present, the development of phosphors for application in phosphor-converted white Light Emitting Diodes (pc-WLEDs) is one of the most relevant areas of research in lighting technologies. In particular, since the introduction of the first commercial pc-WLEDs, a search for new efficient red light-emitting phosphors has been actively ongoing. Indeed, a significant contribution in the red region of the emission spectrum is highly required to obtain warm white light from pc-WLEDs based on the standard technology exploiting the combination of a blue LED chip and a converting yellow YAG:Ce3+-type phosphor [1]. A suitable phosphor should have significant absorption in the blue spectral range and emit in the red (i.e., in the 620 ÷ 650 nm range) [2]. In particular, in most commercial pc-WLEDs, some nitride compounds doped with Eu2+ rare earth ions are used as red phosphors [3]. However, the band width of the emission due to the Eu2+ 5d-4f electronic transition is relatively large, which results in partial emission outside the red region and, as a consequence, a decrease in the luminous efficacy of the red-emitting phosphor [3]. Mn4+-doped luminescent materials with narrow-band emission due to Mn4+ d-d electronic transition have been subjected to intensive studies in the last years as promising red-emitting phosphors under excitation with blue/near UV LEDs.
Among these, in recent years, several red-emitting phosphors doped with Mn4+ have been developed. The attention was mainly oriented to fluoride compounds: here, Mn4+ shows a narrow emission band near 630 nm, which is about optimal for lighting applications. [4]. On the other hand, some drawbacks (e.g., a need for toxic hydrofluoric acid for their synthesis, poor resistance to temperature and humidity) seriously hamper their development. For this reason, at a recent time, studies have also been addressing other candidate hosts such as oxides and oxyfluorides [2,5].
Among oxides, special attention is paid to aluminates in which Mn4+ ions substitute for the Al3+ ions in octahedral sites. Mn4+ and Al3+ have very similar ionic radii but some kind of charge compensation is needed for the stabilization of Mn4+ ions in octahedral sites. The properties of Mn4+ luminescence in different aluminate hosts are described in several review papers (see, e.g., [2]). In particular, Mn4+-doped aluminates whose structures have spinel structured blocks, i.e., the close-packed oxygen layers with Al3+ or Al3+ and Mg2+ in octahedral and tetrahedral sites, were extensively investigated. Such phosphors show luminescence of Mn4+ ions within the red spectral range, namely, SrMgAl10O17 (658 nm) [6], BaMgAl10O17 (660 nm) [7], Sr2MgAl22O36 (659 nm) [8], Ca2Mg2Al28O46 (656 nm) [9], CaMg2Al16O27 (655 nm) [9], Sr4Al14O25 (654 nm) [10], Sr2Al6O11 (652 nm) [11], CaAl12O19 (658 nm) [12], SrAl12O19 (658 nm) [13], and LaMgAl11O19 (663 nm) [14], quite near the edge of eye sensitivity. However, probably the shortest wavelength of Mn4+ luminescence in aluminate hosts is observed for the classical ‘spinel’ MgAl2O4 [15,16,17,18].
Compounds of the spinel group, in particular, the ‘spinel’ itself MgAl2O4, are well-known matrices for developing phosphors. For instance, the MgAl2O4 spinel features a cubic structure (Fd3m space group); in this structure, oxygen anions have a cubic close packing arrangement which creates tetrahedrally and octahedrally coordinated cavities for Mg2+ and Al3+ cations, respectively [19]. In this structure, Mg2+ and Al3+ can exchange their positions in the lattice. Therefore, the distribution of cations can be described with the formula (Mg1-xAlx)[MgxAl2-x]O4, where x is an index expressing the inversion degree. As a result, the spinel crystal structure is disordered, similar to what is observed in solid solutions. Nevertheless, it is generally accepted that this cation inversion allows the fabrication of spinel phosphors where the tetravalent Mn4+ ions substitute for Al3+ ions. This is obtained by introducing an equivalent concentration of Mg2+ ions at the octahedral site. The excess Mg2+ ions compensate for the charge unbalance, still maintaining the stoichiometry with respect to oxygen ions in the spinel crystal lattice [15,16].
In our previous studies it was demonstrated that the synthesis parameters strongly affect the luminescence properties of MgAl2O4 doped with manganese [17,18]. An efficient red-emitting phosphor based on a manganese-doped MgAl2O4 spinel, with a luminescence peak located at ~651 nm, was prepared by means of a low-temperature annealing phase followed by high-temperature annealing, in an oxidizing atmosphere (air). The low-temperature phase allows for efficient entering of Mn4+ in the octahedral sites. The stabilization of Mn4+ ions at octahedral sites is obtained, taking advantage of the presence of additional Mg2+ ions at octahedral sites, which compensate for the charge unbalance of the lattice structure. Conversely, Mn2+ located in the tetrahedral sites produces only pure green (525 nm) emission as observed from a manganese-doped MgAl2O4 spinel synthesized in a reducing CO atmosphere, even if MnO2 is used as a dopant. On the other hand, by using MnO2 or Mn2O3 as dopants, synthesis in neutral argon atmosphere or synthesis without meticulously performed preliminary low-temperature annealing results in the presence of both Mn2+ and Mn4+ in the spinel phosphor as well as the appearance of the other optical centers showing additional emissions; in particular, some broad-band luminescence located at longer wavelengths in the red region is detected. The nature of this far-red broad-band emission of manganese ions in MgAl2O4 phosphors remains unclear, thus further investigations are required.
In the present work, the luminescence properties of a series of ceramic spinel MgAl2O4 phosphors doped with manganese ions and prepared under different synthesis conditions were studied by low-temperature and time-resolved spectroscopy.

2. Materials and Methods

The ceramic samples with the spinel lattice structure were synthesized using MgCO3 (purity of 99.99%) and Al2O3 (99.99%) as well as MnO2 (99.999%) as a source of manganese ions, by means of a high-temperature solid-state reaction. The starting compounds were weighed to obtain the composition Mg1.002Al1.996Mn0.002O4, then mixed in ethanol by using a mortar and a pestle made of agate. The mixture was then dried and uniaxially pressed under about 150 MPa in a stainless-steel die. The resulting pellets were about 2 mm thick and had a diameter of 10 mm. After pressing, the pellets were then placed in a corundum crucible (Thermokeramika, Moscow, Russian Federation). Sample I was subjected to the low-temperature annealing process (consisting of several annealing phases carried out at progressively increasing temperatures, i.e., 500, 600, and 700 °C) and then to the high-temperature annealing process (annealing steps at 1000, 1200, and 1300 °C). Each temperature step lasted for 4 h, in an oxidizing air atmosphere. Conversely, sample II was synthesized without low-temperature annealing. Before each annealing step the tablets were reground and re-pelletized. Moreover, in the last annealing we added to the samples 3 wt% H3BO3 as flux. One more spinel-structured ceramics doped with manganese ions was prepared by using the mixture of sample II, MgO and MgF2 in weighed amounts corresponding to the composition of Mg1.25Al1.75O3.75F0.25 (sample III). This mixture was used to obtain pressed pellets which were subjected to annealing at 1200 °C in an argon atmosphere. The sintered pellets of all samples were finally polished for the following characterization. The thermal treatment used for the three types of samples studied in the present work is presented in Table 1.
The structural-phase composition of the obtained ceramic samples was studied using a Bruker D8 Advance X-ray powder diffractometer (Billerica, MA, USA) with monochromatic CuKα radiation, whose voltage and current were set as 40 kV and 40 mA, respectively. XRD data were recorded in the 2θ range from 10° to 100°, with the continuous scan mode at a speed of 0.2 s per step with a step size of 0.02°. Identification of the synthesized compounds was performed with the software package EVA (Bruker) using the ICDD PDF-2 database. The X-ray diffraction analysis confirmed that all the ceramic samples under test had a lattice structure belonging to the cubic system and corresponding to the spinel structure (Figure 1). The unit cubic cell of the various samples had a lattice parameter in the range of 8.07 ÷ 8.09 Å depending on the synthesis conditions.
Time-resolved photoluminescence (PL) spectroscopy was carried out using a JOBIN-YVON SPEX TRIAX 320 spectrometer (Edison, NJ, USA), equipped with a 300 L/mm grating and using an input slit width of 40 μm, resulting in an overall spectral resolution of 0.6 nm. The temperature T of the samples was set in the interval from T~10 K to T~290 K using a CTI-CRYOGENICS optical cryostat cooled (Mansfield, MA, USA) with a closed cycle refrigerator. For the time-resolved studies (luminescence time decay behavior and time-resolved spectra), pulsed laser excitation was used (wavelength 262 nm, pulse duration 10 ns). For this purpose, a frequency quadrupled Nd:YLF laser was used as the excitation source. To reject the ultraviolet radiation scattered by the sample, a long pass filter (made of Schott GG435 glass) was placed in front of the spectrometer entrance slit. The luminescence time decay behaviors were acquired by means of a photomultiplier tube (Thorn EMI 9816QB, Hayes, UK). A digital sampling scope (Tektronix TDS 680B, Heerenveen, The Netherlands) connected to a PC was used to record the decay curves. The spectral acquisition bandwidth was set between 2 and 4 nm by adjusting the spectrometer exit slit width, depending on the signal intensity. The acquisition of the time-gated spectra was carried out using an optical multichannel detector (detector head EG&G 1420, with a controller EG&G OMA2000, Gaithersburg, MA, USA) with image intensification and time-gating capabilities, featuring a spectral resolution (pixel bandwidth) of 0.51 nm. The delay of the time acquisition gate with respect to the laser pulse excitation and the gate width were set by means of using a Stanford DG535 delay generator (Sunnyvale, CA, USA). No correction was applied for the spectral sensitivity of the detector, which decreases very sharply for wavelengths longer than about 730 nm. This feature can remarkably distort the long-wavelength part of the measured spectra. The working spectral span of the multichannel detector EG&G 1420 is ~320 nm. Accordingly, for obtaining PL spectra in the wide spectral range of 350–850 nm, the measurements were performed for two positions of the monochromator corresponding to two values of the central wavelength of the detector spectral span: 650 and 500 nm. Photo-luminescence excitation (PLE) spectra were measured at 300 K using a spectrofluorometer CM2203 (Solar, Minsk, Republic of Belarus).

3. Results

The ceramic samples of the manganese ion-doped MgAl2O4 spinel labeled as sample I were prepared by using MnO2 as a dopant and using firstly the annealing in air at low temperatures (steps at 500, 600, 700 °C) and after that the annealing at higher temperatures (steps at 1000, 1200, 1300 °C). The use of this elaborate annealing scheme was motivated by the necessity of obtaining the stabilization of the Mn4+ ions at the octahedral site. Indeed, manganese (IV) dioxide begins to decompose at 535 °C to manganese (III) oxide and oxygen [20]. The highest intensity of red emission from Mn4+ in MgAl2O4 was achieved by the final annealing at 1300 °C with the addition of boric acid (H3BO3) as flux.
When excited at 262 nm wavelength, the PL spectrum of sample I, see Figure 2a, is dominated by a relatively broad emission in the red, with a narrow peak at 651 nm. The full-width at half maximum (FWHM) of this peak is ~40 nm at 290 K. According to many previous publications, the observed red emission corresponds well with the luminescence of Mn4+ [15,16,17,18], although a broad-band emission in the blue-green peaking at 430 nm is also observed under 262 nm excitation wavelength. The measurements in the short-wavelength region were performed only for three selected temperatures: 200, 80, and 10 K. The 430 nm band is not detected under Mn4+ intracenter excitation in the near UV/blue spectral region as was shown in previous publications [17,18]. Since this 430 nm emission is observed from all the synthesized samples of MgAl2O4, doped with manganese ions, under 262 nm excitation wavelength, it can tentatively be attributed to some kind of spinel defect-related emission. Accordingly, it is not relevant for practical applications in LEDs, where the excitation occurs in the blue or near UV spectral range.
Under cooling, the PL spectrum of sample I becomes narrower and the short-wavelength wing of the spectrum practically disappears at 10 K, see Figure 2b. Due to the small relative movements of the sample with respect to the laser beam and collection optics during the cooling phase, affecting the signal collection efficiency, it was not possible to compare the absolute luminescence intensity at different temperatures in Figure 2a. For this reason, in Figure 2b, the spectra normalized to the maximum intensity of the peak at 651 nm are presented for the most important red spectral region.
The time-resolved PL spectra from sample I have been measured at a room temperature with time delays between 0 and 1 ms and with a time gate width of 1 ms, see Figure 3. The spinel defect-related broad-band emission disappears already at the shortest delay value 0.1 ms, i.e., this emission has a rather fast decay and does not influence the PL spectrum at longer delays. The decay time of this defect-related emission is estimated to be ~5 µs. Due to this reason, the time-resolved PL spectra were not studied in the spectral range shorter than 500 nm with the position of central wavelength of the detector at 500 nm. The normalized time-resolved PL spectra show that the shape of PL spectrum does not change with the time delay, i.e., luminescence within the whole PL spectrum decays with the same decay law and accordingly corresponds to the decay of the same emitting state.
Decay curves of the red luminescence acquired at different values of temperature after pulsed excitation at 262 nm show non-exponential behavior, see Figure 4, which indicates the presence of some kind of extrinsic quenching of the luminescence, probably due to the energy transfer from Mn4+ ions to some quenching centers originated by defects. The obtained decay curves were fitted by a double exponential decay function. The value of the longer decay component, which can be considered an estimation of intrinsic radiative decay time, is ~0.4 ms at 290 K and practically it does not change with temperature.
On the other hand, in MnO2 doped MgAl2O4 samples subjected to the high temperature annealing only (temperature steps at 1000, 1200, and 1300 °C) but without the low temperature preliminary annealing phase (sample II), the luminescence of Mn4+ ions in the red at 651 nm practically disappears: in the red region the PL spectrum is mainly constituted by the broad-band far-red luminescence band having a maximum between 690 ÷ 710 nm; moreover, an additional green luminescence peak appears at 520 nm, see Figure 5. As it is well-known, the luminescence of manganese ions doped MgAl2O4 in the green is due to the 4T16A1 transition of Mn2+ ions occupying the tetrahedral sites by substituting for Mg2+ ions in the spinel [15,21]. The PLE spectrum of Mn2+, reported in Figure 6, has a characteristic shape featuring four clearly distinguishable bands in the range 350 ÷ 470 nm. One of these bands, typically located near 400 nm (in our case at 430 nm) is rather narrow. The increasing intensity of the green emission at wavelengths shorter than 290 nm is not due to an increase of the Mn2+ luminescence intensity but is caused by the appearance of the broad-band defect-related emission, whose PL spectrum overlaps with that of Mn2+ luminescence. The latter circumstance does not allow studying the spectral properties of Mn2+ luminescence under 262 nm laser excitation separately from those of this defect-related emission.
We note that the PLE spectrum of the emission band in the red with maximum at 651 nm (see Figure 6, red line) has two excitation bands in the blue (peaking at 446 nm) and near UV (peaking at 364 nm). Referring to the scheme of the energy levels of Mn4+ provided by the Tanabe-Sugano (TS) diagram for the d3 electron configuration under an octahedral crystal field (CF) [22], these bands can be attributed to the spin-allowed transitions of Mn4+, namely, 4A24T2 and 4A24T1, respectively. Additionally, the observed excitation band with a peak near 325 nm is due to the O2-—Mn4+ charge-transfer transition. On the other hand, according to Figure 6, the PLE spectrum of broad-band far-red emission coincides with that of Mn4+ luminescence in the range of 300 ÷ 500 nm, i.e., in the region of Mn4+ strong absorption on spin-allowed d-d and fully allowed charge-transfer transitions. Some difference in PLE spectra is observed at wavelengths shorter than 300 nm, in particular, around the excitation laser wavelength and in the range of λ > 500 nm.
When the temperature increases from 10 K to 200 K, the intensities of both the broad-band far-red luminescence as well as Mn2+ green emission decrease but not so strongly as that of Mn4+ luminescence, traces of which can also be seen near 650 nm in the PL spectrum of sample II (Figure 5). The decay time of the broad-band far-red luminescence at 290 K has been estimated to be τ~4.0 ms. Time-resolved PL spectra measured for sample II in the time delay range of 0 ÷ 2.5 ms confirm that the broad-band far-red luminescence has a longer decay time than Mn4+ luminescence and that the latter decays completely at longer time delays (see Figure 7). As in the case of sample I, the spinel defect-related emission disappears from the spectrum of sample II at the shortest time delay of 0.1 ms, and the Mn2+ luminescence band at 520 nm can be better recognized in the time-resolved PL spectra where this luminescence is still observable at long time delays (decay time is in the order of several ms). Again, as in the case of sample I, the time-resolved PL spectra were not studied in the spectral range shorter than 500 nm.
One more peculiarity of magnesium aluminate spinels doped with manganese ions is that the increasing ratio of Mg2+ to Al3+ results in the magnification of the Mn4+ luminescence intensity even if the synthesis of such spinels is carried out in a neutral argon atmosphere. Figure 8 reports the luminescence spectra of manganese ions doped Mg1.25Al1.75O3.75F0.25 spinel (sample III) synthesized in argon by using MgAl2O4:Mn (sample II) as a starting material. The measurements in the short-wavelength region were performed only for three selected temperatures: 200, 80, and 10 K. One can see that in sample III, the Mn4+ luminescence intensity relative to the intensity of the broad-band far-red luminescence is more intense than in sample II, taking into account that the intensities of the broad-band far-red luminescence in these samples have practically the same values. It should also be noted that the ratio of the intensities for the luminescence bands at 651 and 700 nm for sample III is almost constant in the range of temperature from 10 K to 200 K.
In order to demonstrate the difference between the position and relative intensities of the different bands, the PL spectra of all three samples measured at 10 K are presented in one graph in Figure 9.

4. Discussion

The observed low-temperature and time-resolved properties of the red emission band in the PL spectrum of sample I correspond well to those of Mn4+ luminescence from the MgAl2O4 spinel studied earlier at room and higher temperatures [15,16,17,18]. In particular, the narrow feature of the band at 651 nm is the zero-phonon line (ZPL) of the Mn4+ 2E → 4A2 transition, while the bands of the PL spectrum at longer and shorter wavelengths correspond to Stokes and anti-Stokes vibronic sub-bands, respectively. All these PL spectrum features show inhomogeneous broadening due to the cation disorder caused by inversion in the spinel MgAl2O4 structure. In particular, the disorder leads to spectral smearing of vibronic lines. This interpretation is confirmed by the results of the low-temperature measurements which have shown that the main peak at 651 nm considerably narrows with decreasing temperature and the short-wavelength wing of the PL spectrum, i.e., the anti-Stokes vibronic side-band disappears at low temperatures. The time-resolved PL spectra of sample I have shown that all the components of the spectrum are due to a decay from the same initial emitting state, i.e., to the 2E → 4A2 transition in Mn4+ including the vibronic transitions. Thus, it can be concluded that the synthesis procedure applied to sample I provides stabilization of practically all manganese ions in the tetravalent oxidation state.
The obtained properties of Mn2+ green luminescence under 262 nm laser excitation also correspond well to those of Mn2+ in different spinel structures studied earlier although overlapping of the Mn2+ emission spectrum with the spectrum of some spinel defect-related emission does not allow studying these properties in detail. Besides that, the efficiency of Mn2+ luminescence excitation is lower when these samples are excited at 262 nm than in the case of lower-energy excitation.
Some hypotheses can be proposed for the explanation of the nature of the broad-band emission at 675 ÷ 720 nm which is observed in MgAl2O4:Mn phosphor samples. In particular, broad-band luminescence of Mn4+ ions can be observed instead of the narrow-band emission in the case of weak CF strength. The luminescence properties of d3 ions (Mn4+) in octahedral coordination are determined by transitions from the states 2E and 4T2 (i.e., the two lowest-energy excited states); moreover, the energy of the 4T2 level is strongly affected by the CF strength. Indeed, the CF strength determines the position of the 4T2 level with respect to the 2E state as well as the crossover point for these levels in the d3 TS diagram. For this reason, in materials where Mn4+ ions are affected by a strong CF, the narrow-band Mn4+ luminescence associated with the transition 2E → 4A2 is observed, but in materials with crystallographic sites inducing weak CF, the broadband 4T24A2 emission can be observed from Mn4+. However, the latter transition is spin-allowed, i.e., such luminescence should show a relatively fast decay, at least a faster decay than that obtained in our studies for this band (τ ~ 4.0 ms). It could be suggested that the value of the CF strength is near the crossover point for 2E and 4T2 levels, i.e., these levels are thermally coupled, but in this case, the ratio of intensities of the 651 nm peak and this broad band should strongly change with temperature, which is not observed. Besides that, it is unreasonable to expect that in phosphors based on the same host but obtained under different synthesis conditions Mn4+ ions can occupy octahedral sites inducing such different crystal fields: either strong CF (sample I) or weak CF (sample II).
Another hypothesis is that this broad-band emission in the far-red originates from the luminescence of Mn2+ ions located in octahedral sites. This explanation is proposed, for instance in [23] where the broad-band emission in the far red from the defect-rich MgAl2O4 spinel doped with Mn ions is attributed to luminescence of Mn2+ centers in the octahedral sites. If Mn2+ substitutes for Al3+ in the octahedral sites the charge compensation can be reached through inversion without introducing other ions. According to the TS energy-level diagram for d5 electron configuration, luminescence of Mn2+ ions can be observed in significantly different spectral ranges depending on the strength of CF affecting these ions. In particular, luminescence in the green spectral range is expected for Mn2+ ions entering the tetrahedral sites whereas red and even NIR luminescence can be observed from Mn2+ occupying octahedral sites. It should be noted here that in the case of the d5 electron configuration, the TS diagrams for octahedral and tetrahedral coordination are identical by taking into account that the CF strength is stronger for octahedral sites than for tetrahedral ones.
As all the Mn2+ absorption transitions are spin-forbidden; the corresponding bands in the PLE spectrum of Mn2+ can be concealed by the wider and more intense bands caused by spin-allowed transitions within Mn4+. In this case the PLE spectra of such possible far-red Mn2+ luminescence and the red Mn4+ luminescence in the blue and near UV regions can coincide if one considers that the far-red luminescence from Mn2+ is excited by the energy transfer from Mn4+ ions to Mn2+ ions. This similarity of PLE spectra is indeed observed in the experiment for the broad-band far-red luminescence and the red Mn4+ luminescence. On the other hand, the PLE spectrum of 520 nm emission corresponding to the luminescence of Mn2+ entered into the spinel tetrahedral sites is well reproduced with all its specific features, see Figure 6. Therefore, it is unclear why the respective features are not seen in the PLE spectrum of the observed far-red luminescence if this far-red luminescence originates from the Mn2+ ions entered into octahedral sites. Besides that, one could hardly suggest that Mn2+ ions having an ionic radius of 0.83 Å in octahedral coordination can substitute for Al3+ ions having an ionic radius of 0.535 Å in the same one [24], i.e., the ionic radii of these octahedrally coordinated ions differ by more than 50%, and according to standard crystallographic rules, such substitution is highly unlikely by taking into account that the ratio of Mg2+ and Mn2+ ionic radii in tetrahedral coordination is 0.57/0.66 Å. In other words, it is most likely that in the spinel MgAl2O4 structure, the Mn2+ ions occupy the tetrahedral sites but not the octahedral ones.
Generally speaking, it can be assumed that in compounds doped with Mn ions, the octahedral sites contain both Al3+ as well as Mn3+ ions, for which charge compensation is not required. Information on the luminescence behavior of Mn3+ is very limited [25,26]. In general, it is expected that in the majority of the hosts, the Mn3+ luminescence is quenched due to the large Jahn-Teller splitting of Mn3+ energy levels. Considering the TS diagram for d4 ions, the broad-band luminescence of Mn3+ entering into the octahedral site can be due to either the spin-allowed 5T25E transition which has typically rather short decay time (of the order of tens μs [26]) or the slower spin-forbidden 1T25E transition, depending on the CF strength. The Mn3+ broad-band absorption is due to the strong spin-allowed 5E → 5T2 transition, typically located at 500 ÷ 550 nm, i.e., just in the spectral range where some additional band in the PLE spectrum of emission at 689 nm is observed for sample II, see Figure 6. In the spectral range 325 ÷ 500 nm, absorption by Mn4+ ions dominates and the shape of the PLE spectrum of the broad-band far-red luminescence almost coincides with that of Mn4+ luminescence monitored at 651 nm, which can be simply explained by the overlapping of emission bands of these two kinds of luminescence centers but can be partly due to the presence of the energy transfer from Mn4+ to Mn3+. However, at wavelengths shorter than ~325 nm, the broad-band far-red luminescence is excited more efficiently, which can be due to some kind of O 2p-Mn 3d charge-transfer transition. The decay time of this broad-band far-red luminescence is rather long (i.e., τ ~ 4.0 ms) and so this luminescence can be ascribed to the Mn3+ spin-forbidden 1T25E transition. In the presence of charge compensation by extra amounts of Mg2+ ions, the efficient conversion of Mn3+ to Mn4+ takes place, resulting in the appearance of Mn4+ luminescence in sample III and in the complete disappearance of this broad-band far-red luminescence in sample I synthesized under oxidation conditions.

5. Conclusions

The obtained low-temperature and time-resolved features of the red emission band characterizing the PL spectrum of the Mn4+-doped MgAl2O4 sample confirmed the generally accepted model of this luminescence as caused by the Mn4+ 2E → 4A2 transitions, including the ZPL located at 651 nm and the Stokes and anti-Stokes vibronic side-bands, which are broadened by the cation disorder caused by inversion in the spinel crystal structure. In this work it has been shown that the special multi-step annealing procedure applied for the solid-state synthesis of a MgAl2O4:Mn4+ phosphor provides the stabilization of practically all manganese ions introduced into the phosphor in the tetravalent state. Furthermore, red-emitting MgAl2O4:Mn4+ phosphor synthesized by this method demonstrates good color characteristics (CIE 1931 color coordinates are x = 0.72; y = 0.28 [18]); it can be therefore considered as promising red phosphor for the realization of warm pc-WLEDs. The broad-band far-red luminescence observed from MgAl2O4 phosphors containing manganese ions non-stabilized in the tetravalent state has been attributed to spin-forbidden 1T25E transitions of Mn3+ ions substituting for Al3+ ions in the octahedral sites in the spinel structure. Such MgAl2O4:Mn3+ phosphors can be used for different lighting applications, among them those related to the agricultural lighting [27].

Author Contributions

Conceptualization, N.K., M.B. and V.M.; methodology, N.K.; formal analysis, V.M.; investigation, G.T., A.P., B.P. and M.V.; data curation, A.P.; writing—original draft preparation, V.M.; writing—review and editing, V.M., N.K., G.T., A.P., B.P. and M.V.; project administration, M.B.; funding acquisition, A.P. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Russian Science Foundation (RSF), grant number 18-13-00407 for the synthesis, X-ray analysis, and PLE spectral studies and by Ente Cassa di Risparmio di Firenze (IFAC-CNR, Rif. 2018.1124) for the time-resolved and temperature-resolved PL spectral studies.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data is contained within the article. Further numerical data (e.g., spectra shown in Figures) are available on request from the corresponding Author.

Acknowledgments

This work was supported by the Russian Science Foundation (RSF), grant number 18-13-00407 “Synthesis of the new red phosphors based on fluorine-containing materials, doped with optically active manganese ions, for warm white light emitting diodes” and by Ente Cassa di Risparmio di Firenze (Rif. 2018.1124), project “Sviluppo e fabbricazione di fosfori, non contenenti elementi chimici appartenenti al gruppo delle Terre Rare, da usare per costruire dispositivi LED con emissione di luce bianca”.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Zhou, Q.; Dolgov, L.; Srivastava, A.M.; Zhou, L.; Wang, Z.; Shi, J.; Dramićanin, M.D.; Brik, M.G.; Wu, M. Mn2+ and Mn4+ red phosphors: Synthesis, luminescence and applications in WLEDs. A review. J. Mater. Chem. C 2018, 6, 2652–2671. [Google Scholar] [CrossRef]
  2. Adachi, S. Review—Mn4+-activated red and deep red-emitting phosphors. ECS J. Solid State Sci. Technol. 2020, 9, 016001. [Google Scholar] [CrossRef]
  3. Lin, C.; Chen, W.T.; Liu, R. Phosphors for White LEDs. In Handbook of Advanced Lighting Technology; Karlicek, R., Sun, C.C., Zissis, G., Ma, R., Eds.; Springer: Cham, Switzerland, 2017. [Google Scholar] [CrossRef]
  4. Adachi, S. Photoluminescence spectra and modeling analyses of Mn4+-activated fluoride phosphors: A review. J. Lumin. 2018, 197, 119–130. [Google Scholar] [CrossRef]
  5. Khaidukov, N.; Brekhovskikh, M.; Toci, G.; Patrizi, B.; Vannini, M.; Pirri, A.; Makhov, V. Time-and temperature-resolved luminescence spectroscopy of LiAl4O6F: Mn red phosphors. J. Lumin. 2019, 216, 116754. [Google Scholar] [CrossRef]
  6. Cao, R.; Xue, H.; Yu, X.; Xiao, F.; Wu, D.; Zhang, F. Luminescence properties and synthesis of SrMgAl10O17:Mn4+ red phosphor for white light-emitting diodes. J. Nanosci. Nanotechnol. 2016, 16, 3489–3493. [Google Scholar] [CrossRef] [PubMed]
  7. Wang, B.; Lin, H.; Huang, F.; Xu, J.; Chen, H.; Lin, Z.; Wang, Y. Non-rare-earth BaMgAl10–2xO17:xMn4+,xMg2+: A narrow-band red phosphor for use as a high-power warm w-LED. Chem. Mater. 2016, 28, 3515–3524. [Google Scholar] [CrossRef]
  8. Zhang, H.; Zhang, H.; Zhuang, J.; Dong, H.; Zhu, Y.; Ye, X.; Liu, Y.; Lei, B. Effect of H3BO3 flux on the morphology and optical properties of Sr2MgAl22O36:Mn4+ red phosphors for agricultural light conversion films. Ceram. Int. 2016, 42, 13011–13017. [Google Scholar] [CrossRef]
  9. Sasaki, T.; Fukushima, J.; Hayashi, Y.; Takizawa, H. Synthesis and photoluminescence properties of Mn4+-doped magnetoplumbite-related aluminate X-type Ca2Mg2Al28O46 and W-type CaMg2Al16O27 red phosphors. Ceram. Int. 2017, 43, 7147–7152. [Google Scholar] [CrossRef]
  10. Meng, L.; Liang, L.; Wen, Y. An efficient red phosphor Li+, Mn4+ co-doped Sr4Al14O25 for warm white light emitting diodes. Sci. Adv. Mater. 2017, 9, 456–459. [Google Scholar] [CrossRef]
  11. Sasaki, T.; Fukushima, J.; Hayashi, Y.; Takizawa, H. Synthesis and photoluminescence properties of a novel Sr2Al6O11:Mn4+ red phosphor prepared with a B2O3 flux. J. Lumin. 2018, 194, 446–451. [Google Scholar] [CrossRef]
  12. Zhao, Y.; Shi, L.; Han, Y.-J.; Li, H.-H.; Ji, Z.-X.; Zhang, Z.-W. Luminescent properties of Zn2+- doped CaAl12O19:Mn4+ deep-red phosphor for indoor plant cultivation. Ceram. Int. 2019, 45, 8265–8270. [Google Scholar] [CrossRef]
  13. Wang, X.; Li, P.; Brik, M.G.; Li, X.; Li, L.; Peng, M. Thermal quenching of Mn4+ luminescence in SrAl12O19:Mn4+. J. Lumin. 2019, 206, 84–90. [Google Scholar] [CrossRef]
  14. Li, X.; Chen, Z.; Wang, B.; Liang, R.; Li, Y.; Kang, L.; Liu, P. Effects of impurity doping on the luminescence performance of Mn4+-doped aluminates with the magnetoplumbite-type structure for plant cultivation. Materials 2019, 12, 86. [Google Scholar] [CrossRef] [Green Version]
  15. Wakui, Y.; Shan, Y.J.; Tezuka, K.; Imoto, H.; Ando, M. Crystal-site engineering approach for preparation of MgB2O4:Mn2+, Mn4+ (B = Al, Ga) phosphors: Control of green/red luminescence properties. Mat. Res. Bull. 2017, 90, 51–58. [Google Scholar] [CrossRef]
  16. Zhang, Y.; Liu, Y.; Yang, L.; Hu, S.; Wang, Z.; Jing, Z.; Nian, H.; Liu, B.; Zhou, G.; Wang, S. Preparation and luminescence properties of thermally stable Mn4+ doped spinel red-emitting ceramic phosphors. J. Lumin. 2020, 220, 117016. [Google Scholar] [CrossRef]
  17. Khaidukov, N.M.; Brekhovskikh, M.N.; Kirikova, N.Y.; Kondratyuk, V.A.; Makhov, V.N. Luminescence of MgAl2O4 and ZnAl2O4 spinel ceramics containing some 3d ions. Ceram. Int. 2020, 46, 21351–21359. [Google Scholar] [CrossRef]
  18. Khaidukov, N.M.; Brekhovskikh, M.N.; Kirikova, N.Y.; Kondratyuk, V.A.; Makhov, V.N. Luminescent properties of spinels doped with manganese ions. Russ. J. Inorg. Chem. 2020, 65, 1135–1141. [Google Scholar] [CrossRef]
  19. Bosi, F.; Biagioni, C.; Pasero, M. Nomenclature and classification of the spinel supergroup. Eur. J. Mineral. 2019, 3, 183–192. [Google Scholar] [CrossRef] [Green Version]
  20. Cahiez, G.; Alami, M.; Taylor, R.J.K.; Reid, M.; Foot, J.S.; Fader, L.; Sikervar, V.; Pabba, J. Manganese Dioxide. In Encyclopedia of Reagents for Organic Synthesis; Wiley: Hoboken, NJ, USA, 2017. [Google Scholar] [CrossRef]
  21. Song, E.H.; Zhou, Y.Y.; Wei, Y.; Han, X.X.; Tao, Z.R.; Qiu, R.L.; Xia, Z.G.; Zhang, Q.Y. A thermally stable narrow-band green-emitting phosphor MgAl2O4: Mn2+ for wide color gamut backlight display application. J. Mater. Chem. C 2019, 7, 8192–8198. [Google Scholar] [CrossRef]
  22. Tanabe, Y.; Sugano, S. On the absorption spectra of complex ions II. J. Phys. Soc. Jpn. 1954, 9, 776–779. [Google Scholar] [CrossRef]
  23. Mohler, R.L.; White, W.B. Influence of structural order on the luminescence of oxide spinels: Manganese activated spinels. Mater. Res. Bull. 1994, 29, 1109–1116. [Google Scholar] [CrossRef]
  24. Shannon, R.D. Revised effective ionic radii and systematic studies of interatomic distances in halides and chalcogenides. Acta Cryst. A 1976, 32, 751–767. [Google Scholar] [CrossRef]
  25. van Die, A.; Leenaers, A.C.H.I.; van der Weg, W.F.; Blasse, G. A search for luminescence of the trivalent manganese ion in solid aluminates. Mater. Res. Bull. 1987, 22, 781–787. [Google Scholar] [CrossRef] [Green Version]
  26. Kück, S.; Hartung, S.; Hurling, S.; Petermann, K.; Huber, G. Optical transitions in Mn3+-doped garnets. Phys. Rev. B 1998, 57, 2203–2216. [Google Scholar] [CrossRef]
  27. Piovene, C.; Orsini, F.; Bosi, S.; Sanoubar, R.; Bregola, V.; Dinelli, G.; Gianquinto, G. Optimal red:blue ratio in led lighting for nutraceutical indoor horticulture. Sci. Hortic. 2015, 193, 202–208. [Google Scholar] [CrossRef]
Figure 1. X-ray diffraction pattern of the MgAl2O4 spinel doped with 0.2 mol% MnO2 (sample I). The labels on the peaks are the Miller indices of the corresponding lattice planes.
Figure 1. X-ray diffraction pattern of the MgAl2O4 spinel doped with 0.2 mol% MnO2 (sample I). The labels on the peaks are the Miller indices of the corresponding lattice planes.
Materials 14 00420 g001
Figure 2. (a) PL spectra of MgAl2O4:Mn4+ (sample I) measured at different temperatures: 290, 200, 150, 100, 80, 50, 40, 30, 20, and 10 K; (b) PL spectra of MgAl2O4:Mn4+ (sample I), normalized at their peak value, for the red spectral region measured at different temperatures.
Figure 2. (a) PL spectra of MgAl2O4:Mn4+ (sample I) measured at different temperatures: 290, 200, 150, 100, 80, 50, 40, 30, 20, and 10 K; (b) PL spectra of MgAl2O4:Mn4+ (sample I), normalized at their peak value, for the red spectral region measured at different temperatures.
Materials 14 00420 g002
Figure 3. (a) Time-resolved PL spectra of the spinel MgAl2O4 ceramics doped with 0.2 mol% MnO2 (sample I) measured at 290 K with different delays (0–1 ms) between the laser pulse and the time gate (Δt = 1 ms); (b) the same spectra normalized to the peak intensity.
Figure 3. (a) Time-resolved PL spectra of the spinel MgAl2O4 ceramics doped with 0.2 mol% MnO2 (sample I) measured at 290 K with different delays (0–1 ms) between the laser pulse and the time gate (Δt = 1 ms); (b) the same spectra normalized to the peak intensity.
Materials 14 00420 g003
Figure 4. Normalized decay curves of red (λem = 663 nm) luminescence recorded from sample I, MgAl2O4 doped with 0.2 mol% MnO2, at T from 10 K to 200 K.
Figure 4. Normalized decay curves of red (λem = 663 nm) luminescence recorded from sample I, MgAl2O4 doped with 0.2 mol% MnO2, at T from 10 K to 200 K.
Materials 14 00420 g004
Figure 5. PL spectra of sample II, i.e., MgAl2O4:Mn, measured at different temperatures: 200 K, 100 K, and 10 K.
Figure 5. PL spectra of sample II, i.e., MgAl2O4:Mn, measured at different temperatures: 200 K, 100 K, and 10 K.
Materials 14 00420 g005
Figure 6. PLE spectra of sample II, i.e., MgAl2O4:Mn, measured at 290 K by monitoring different emission wavelengths.
Figure 6. PLE spectra of sample II, i.e., MgAl2O4:Mn, measured at 290 K by monitoring different emission wavelengths.
Materials 14 00420 g006
Figure 7. Time-resolved PL spectra of MgAl2O4:Mn (sample II) measured at 290 K at different delays (0 ÷ 2.5 ms) of the time gate (Δt = 1 ms) with respect to the laser pulse.
Figure 7. Time-resolved PL spectra of MgAl2O4:Mn (sample II) measured at 290 K at different delays (0 ÷ 2.5 ms) of the time gate (Δt = 1 ms) with respect to the laser pulse.
Materials 14 00420 g007
Figure 8. Normalized PL spectra of manganese ions doped Mg1.25Al1.75O3.75F0.25 (sample III), measured at different temperatures.
Figure 8. Normalized PL spectra of manganese ions doped Mg1.25Al1.75O3.75F0.25 (sample III), measured at different temperatures.
Materials 14 00420 g008
Figure 9. Normalized PL spectra of samples I, II, and III measured at 10 K.
Figure 9. Normalized PL spectra of samples I, II, and III measured at 10 K.
Materials 14 00420 g009
Table 1. Experimental conditions applied for the synthesis of ceramic spinels containing manganese ions.
Table 1. Experimental conditions applied for the synthesis of ceramic spinels containing manganese ions.
Sample LabelHostAtmosphereAnnealing Temperature (°C)
Sample IMgAl2O4air500, 600, 700 and 1000, 1200, 1300
Sample IIMgAl2O4air1000, 1200, 1300
Sample IIIMg1.25Al1.75O3.75F0.25argon1200
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Khaidukov, N.; Pirri, A.; Brekhovskikh, M.; Toci, G.; Vannini, M.; Patrizi, B.; Makhov, V. Time- and Temperature-Dependent Luminescence of Manganese Ions in Ceramic Magnesium Aluminum Spinels. Materials 2021, 14, 420. https://doi.org/10.3390/ma14020420

AMA Style

Khaidukov N, Pirri A, Brekhovskikh M, Toci G, Vannini M, Patrizi B, Makhov V. Time- and Temperature-Dependent Luminescence of Manganese Ions in Ceramic Magnesium Aluminum Spinels. Materials. 2021; 14(2):420. https://doi.org/10.3390/ma14020420

Chicago/Turabian Style

Khaidukov, Nicholas, Angela Pirri, Maria Brekhovskikh, Guido Toci, Matteo Vannini, Barbara Patrizi, and Vladimir Makhov. 2021. "Time- and Temperature-Dependent Luminescence of Manganese Ions in Ceramic Magnesium Aluminum Spinels" Materials 14, no. 2: 420. https://doi.org/10.3390/ma14020420

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