Next Article in Journal
Corrosion-Resistant Plug Materials for Geothermal Well Fluid Control
Previous Article in Journal
Laboratory Study and Field Validation of the Performance of Salt-Storage Asphalt Mixtures
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Ho-SiAlON Ceramics as Green Phosphors under Ultra-Violet Excitations

1
Research Center for Green Advanced Materials, Sun Moon University, Chungnam 31460, Korea
2
Department of Fusion Science and Technology, Sun Moon University, Chungnam 31460, Korea
3
Department of Environment and Bio-Chemical Engineering, Sun Moon University, Chungnam 31460, Korea
*
Authors to whom correspondence should be addressed.
Materials 2022, 15(19), 6715; https://doi.org/10.3390/ma15196715
Submission received: 22 July 2022 / Revised: 20 September 2022 / Accepted: 23 September 2022 / Published: 27 September 2022
(This article belongs to the Section Advanced and Functional Ceramics and Glasses)

Abstract

:
In most inorganic phosphors, increasing the concentration of activators inevitably causes the concentration quenching effect, resulting in reduced emission intensity at a high level of activator doping and the conventional practice is to limit the activator concentration to avoid the quenching. In contrast, SiAlON ceramics preserve their chemical composition over a very wide range of doping of activator ions, which favors the adjustment and optimization of the luminescence properties avoiding concentration quenching. Here, we investigate the photoluminescence properties of Ho-doped SiAlON (Ho-SiAlON) ceramics phosphors prepared by the hot-press method. Ho-SiAlON ceramics show strong green visible (554 nm) as well as infrared (2046 nm) broadband downshifting emissions under 348 nm excitation. It is shown that there is no concentration quenching, even at a very high level of Ho doping. The emission intensity of the 554 nm band increased two-fold when the Ho concentration is doubled. The results show that the Ho-SiAlON ceramics can be useful for efficient green phosphors.

1. Introduction

SiAlON ceramics are known to have an incomparable balance of mechanical and thermal properties [1,2] that enable them to resist extreme thermal shock and keep mechanical integrity in temperatures as high as 1673 K (1400 °C). For this reason, SiAlON ceramics are preferred for use in refractory applications over traditional ceramics, such as zirconia (ZrO2) and alumina (Al2O3). SiAlON ceramics are generally prepared at high temperatures using the solid-state reaction technique. The precursor is an appropriate mixture of α-Si3N4, AlN, Al2O3, and a small amount of oxides of Ca, Mg, or lanthanides (Ln3+) as a sintering additive [3]. The Ln3+ ion that facilitates the sintering of α-Si3N4 [4] helps to keep the charge neutrality in the α-SiAlON crystal during the preferential substitutions of (Si, Al) and (N, O) in the lattice [5]. Moreover, Ln3+ ions also impart the α-SiAlON ceramics with efficient photoluminescence [6,7]. α-SiAlON ceramics have been reported as efficient inorganic remote phosphors for white light emitting diodes (WLEDs) [8,9].
Generally, an increase in the activator concentration should be accompanied by an increase in the intensity of the emitted light due to increased absorption efficiency of the excitation energy. However, in most inorganic phosphors, increasing the concentration of activators inevitably decreases the emission intensity [10,11,12,13,14,15], a phenomenon called concentration quenching. Concentration quenching of the emission is attributed to the increased probability of energy migration between the luminescent centers as the distance between the luminescent centers becomes shorter with increasing activator concentration [16]. Conventional practice is to limit the activator concentration to avoid concentration quenching. In this regard, SiAlON ceramics preserve their chemical composition over a very wide range of doping of activator ions, which favors the adjustment and optimization of the luminescence properties without suffering from concentration quenching.
In addition to the high level of doping ability of activators, the thermal stability of the host is also very important for stable optical performance. This is because luminescence stability is closely associated with the structural stability of the phosphor. The stronger the structural rigidity, the better the thermal stability of the host, and hence, stable the luminescence. Many of the efficient glass-based conventional phosphors [17] have low thermal stability. In such phosphors, the lattice thermal expansion due to heat may change the local environment of the crystal site occupied by the Ln3+ ion [18,19], ultimately causing thermal quenching of the emission bands. In this regard, phosphors with a very low thermal expansion coefficient are preferred for stable luminescence. SiAlON ceramics are known to have a very low thermal expansion coefficient [1], and therefore, have high thermal stability, even at extreme temperatures. Therefore, studies on the Ln3+-doped SiAlON phosphors have a considerable research value. The holmium ion (Ho3+) can be efficiently excited in the ultraviolet (UV) and blue spectral range giving strong green luminescence due to the 4f-4f transition of the Ho3+ ion. Hence, Ho-SiAlON ceramics can be potential green phosphors.
In this report, we synthesize Ho-SiAlON ceramics by the hot-press technique and investigate their photoluminescence properties under UV excitations. The ceramic samples are characterized for phase and morphology using XRD and SEM, respectively. The valence state of the dopant has been confirmed by the XPS analysis. We study Ho concentration-dependent photoluminescence of Ho-SiAlON ceramics and find that these ceramics have a high level of doping tolerance of Ho ions and do not suffer from concentration quenching. It will be shown that the ceramics can be useful for green phosphors.

2. Experimental Section

A series of Ho-SiAlON ceramics were obtained by the hot-press sintering method. Sample compositions were designed according to HoxSi12−m−n Alm + nOnN16−n; x = m/v, where x is solubility and v is the valency of the metal cation Ho [3]. The starting m and n parameters and precursors composition are given in Table 1. Four different sets of m and n parameters were chosen as listed in Table 1. α-Si3N4 (UBE Co., Tokyo, Japan), AlN (Tokuyama Co., Tokyo, Japan), Al2O3, and Ho2O3 (High purity chemicals Co LTD., Saitama, Japan) were used as starting materials. The samples were hot-press sintered at 1850 °C, as described elsewhere [20,21]. The samples were re-sized and polished for various characterizations.
The X-ray diffraction (XRD) spectra were obtained using Cu Kα radiation (Rigaku D/Max 2200HR diffractometer, Akishima, Tokyo, Japan) at a scanning rate of 1° per minute in the 2θ range of 10° to 90°. The X-ray photoelectron spectra (XPS) were recorded by Versa Probe spectrometer (PHI 5000, Ulvac-PHI, Kanagawa, Japan) with monochromated Al Kα (1486.6 eV) as an X-ray source. The transmission electron microscopy (TEM) images and energy-dispersive spectra (EDS) were recorded by using TEM (JEM-4010, JEOL, Tokyo, Japan). Absorption spectra in the 200–2500 nm range were measured by a V-570 spectrophotometer (Jasco International Co. LTD., Tokyo, Japan). The photoluminescence spectra were obtained using an FLS980 spectrometer (Edinburgh Instruments, Edinburgh, UK). A 400 Watt Xenon lamp was used for UV excitation in CW mode and a microsecond flash lamp (µf2, Edinburgh Instruments, Edinburgh, UK) was used for pulsed excitation.

3. Results and Discussion

3.1. Phase and Microstructure Analysis

Figure 1 displays the XRD spectra of all the sintered samples and shows that the Ho-SiAlON samples have similar XRD patterns. The peaks have been indexed with reference to the XRD spectra of the Ca-α-SiAlON (JCPDS 33-261). The chosen m and n parameters are within the solubility limits of α-SiAlON solid solutions [22,23]. However, there are a few minor peaks of the SiAlON polytypoid phase [24]. The samples show that there is a gradual shift of the (210) peak (Figure 1b) towards a smaller angle with the m parameter increasing from m = 1.0 to m = 2.0. With an increasing m value, the content of Ho increases (Table 1) in the sample. Hence, the shifting of the (210) plane with increasing m value can be due to the increase in the unit cell size [23].
Figure 2 shows the SEM images of the fracture surface of all the samples. A fully dense α-SiAlON microstructure without pores is evident from the micrograph in Figure 2. It can be noticed that the grain morphology changes with increasing Ho concentration. Samples H10 and H11 are composed of isometric polyhedral grains, while more elongated grains can be seen in samples H15 and H20. The fracture mode also appears to be different in different samples. The transverse grain fracture and grain pull-out can be seen in samples H10 and H11, while lateral grain sliding can be observed in samples H15 and H20. Additionally, the TEM micrograph of sample H15 in Figure 3 shows the grains of the α-SiAlON phase with grain sizes ranging from sub-micron to a few micrometers. Moreover, the EDS spectra of different points across the sample show the presence of all the elements. Figure 4 shows the high-angle annular dark-field (HAADF) TEM image and corresponding elemental mappings of H15 sample. Homogeneous distribution of Si, Al, O, and N across the matrix of Ho-α-SiAlON can be seen. However, a relatively higher concentration of Ho ions at the triple junction (point 2 and point 5) can be observed in Figure 4f,g. The higher concentration of Ho ions at the triple junction is due to the segregation of the Ho ions, which are not incorporated into the α-SiAlON lattice during sintering.

3.2. XPS Analysis

XPS has been extensively used for the characterization of Si3N4-based thin films [25,26], powders [27], and sintered materials [28]. However, XPS studies on the hot-pressed α-SiAlON ceramics are limited [29]. The XPS survey spectra and high-resolution elemental spectra of the H15 sample are shown in Figure 5. The peaks corresponding to each constituent element have been fitted and the results are presented in Table 2. A carbon peak at a binding energy of 284.83 eV observed in the survey spectra (Figure 5a) was used for the calibration.
In Figure 5b, the Si 2p peak has two sub-peaks. Since the binding energy of the photoelectrons increases with the increase in electronegativity of the nearest neighbors of an atom, the Si 2p line can be expected to be at high binding energy due to the high electronegativity of oxygen atoms with respect to that of nitrogen atoms [30]. Therefore, the dominant sub-peak of lower binding energy corresponds to the Si-N bond in the Si3N4 matrix [31], while the less intense sub-peak of higher binding energy corresponds to the Si-O bond in SiO2 [32] that exist in the grain boundary glassy phase of the SiAlON ceramics [33]. In Figure 5c, the Al 2p scan has two sub-peaks. The peak with lower binding energy corresponds to the Al-N bond that originates from AlN-rich phases in SiAlON ceramics and the peak with higher binding energy can be assigned to the Al-O bond [34]. The Al-O bond appears as a result of the partial Al(Si)-O(N) substitution in the Si3N4 matrix during the formation of the α-SiAlON lattice [3]. In Figure 5d, the O1s scan consists of two peaks. The intense peak at lower binding energy can be assigned to the O-Al bond that results due to Al(Si)-O(N) substitution as discussed earlier. The less intense peak at higher binding energy is due to the O-Si bond of the intergranular glassy phases. The N1s peak consists of three sub-peaks as shown in Figure 5e. The weak sub-peak at the lowest binding energy corresponds to the N-Al, while the intense central peak corresponds to the N-Si bonds in the α-SiAlON lattice because the N1s binding energy for AlN is less than that for Si3N4 [35]. In the α-SiAlON crystal, the Ln3+ dopant (Ho3+) occupies the seven-fold coordination site [5,21] for the charge compensation which arises due to the partial Al (Si)-O (N) substitution in the formation of the SiAlON lattice [3]. Hence, it is suggested that the broad sub-peak with the highest binding energy in Figure 5e corresponds to the N-Ho bond. The XPS spectrum of Ho4d in Figure 5f is the characteristic of the trivalent Ho ion [36]. Hence, it can be inferred that the Ho ion exists in a 3+ valance state in the α-SiAlON lattice.

3.3. Absorption Spectra:

The absorption spectra of all the Ho-SiAlON samples are shown in Figure 6 in the range of 250–2500 nm. All the samples show the characteristic absorption peaks of the Ho3+ ion. The absorption peaks at 340, 348, 364, 390, 422, 460, 488, 543, 650, 1148, and 1945 nm are assigned, respectively, to the transitions to the excited states 3F2, 5G3, 3H6, 5G4 + 3K7, 5G5, 5F1 + 5G6, 5F2,3, 5F4 + 5S2, 5F5, 5I6, and 5I7 from the ground state 5I8 of Ho3+ ion [37]. Below 320 nm, an ultraviolet absorption edge exists in all the samples, which can be assigned to the charge transfer within the Si(Al)-N(O) network in the Ho-SiAlON matrix [38].

3.4. Downshifting Photoluminescence in Ho-SiAlON Ceramics

Emission spectra of Ho-SiAlON samples in the visible spectral range under 348 nm continuous wave (CW) as well as pulsed excitation are shown in Figure 5a,b, respectively. The 348 nm excitation corresponds to the 5I85G3 absorption transition of Ho3+ ions (Figure 10). There are three sharp and strong emission bands at 554 nm (green), 664 nm (red), and 760 nm (near-infrared), which are assigned to the transitions 5F4, 5S25I8, 5F55I8, and 5F4, 5S25I7, respectively of the Ho3+ ion [39]. A weak band at 590 nm is assigned to the 5G45I6 transition. Intensities of the 554 and 760 nm emission bands increase with the increase in Ho concentration. As shown in Figure 7a, a nearly two-fold increase in the intensity of the 554 nm emission band is observed for the H20 sample (Ho-concentration 18.068 wt.%) as compared to the H10 sample (Ho-concentration 9.791 wt.%). The existence of these emission bands is an indication that the Ho3+ ions have been incorporated into the SiAlON lattice.
The emission spectra under CW (Figure 7a) and pulsed excitation (Figure 7b) are similar, but a slight change in emission intensity can be observed. The intensity modulation under CW and pulsed excitations can occur, and therefore, this results in slightly different emission colors, as can be seen from the CIE 1931 chromaticity diagram in Figure 8a,b. The CIE color coordinates presented in Table 3 under both the CW and pulsed excitations are in the green spectral region. Hence, Ho-SiAlON ceramics can be useful for green phosphors. A particular advantage of using Ho-SiAlON as green phosphor material is that there is no reduction of the emission intensity due to the concentration quenching, even at a very high level of Ho doping as shown above. On contrary, many other host materials, such as LaMgAl11O19 [40], Y2O3 [41], NaYF4 [13], K2SiF6 [12], etc., show a decrease in emission intensity due to concentration quenching at a high level of doping of the activators. An additional advantage of Ho-SiAlON ceramics as phosphors is that SiAlON ceramics have a very high laser damage threshold and high-temperature spectral stability [21,42] that make them suitable for high-power green emitting remote phosphors in extreme conditions [8,9].
Under 348 nm CW excitation, a broad mid-infrared emission band from 1850 to 2150 nm with FWHM of 186 nm is observed as shown in Figure 9, which can be assigned to 5I75I8 transition of Ho3+. The multiple peak profile of the emission band originates due to the rich Stark splitting of the 5I7 and 5I8 emitting states. A weak emission at 1196 nm corresponds to 5I65I8, whose Ho-concentration dependence is not evident, but there is a steady increase in the 2046 nm emission intensity with the increases in Ho concentration. The 2046 nm broadband emission of Ho3+ has a wide range of applications, such as in infrared lasers [43].
Downshifting emissions from Ho3+ under UV excitations have been reported for other crystal systems [44,45,46]. The observed downshifting emission properties of Ho-SiAlON under 348 nm excitation can be explained using the energy level diagram in Figure 10. Upon excitation at 348 nm, Ho3+ ions are first excited to the 5G3 manifold, and then via successive non-radiative relaxations, 5G4, 5F2,3, 5F4, and 5S2 manifolds are populated. A weak radiative transition at 590 nm appears from transition 5G45I6, which is followed by the 5I65I8 transition to produce another weak infrared emission at 1196 nm. However, strong red emissions at 664 and 760 nm are produced by the radiative transitions 5F2,35I7 and 5F4, 5S25I7, respectively. Both of these radiative transitions are followed by 5I75I8 to produce an intense infrared emission at 2046 nm. There are also other decay pathways of 5F4 and 5S2 manifolds to the ground state 5I8. Some of the Ho3+ ions in 5F4, 5S2 manifolds can directly undergo radiative decay to the ground state 5I8 to produce strong green emission at 554 nm, while other Ho3+ ions decay via the intermediated non-radiative relaxation 5F4, 5S25F5 to the ground state 5I8 via the decay to produce a strong red emission at 664 nm.

4. Conclusions

A series of Ho-SiAlON ceramics were prepared by the hot-press sintering method, and their photoluminescence properties were investigated under UV excitations. A fully dense α-SiAlON phase of all the sintered samples has been verified using XRD and TEM analysis. Moreover, the XPS analysis confirms that the Ho ions exist in a trivalent state. Ho-SiAlON ceramics show strong downshifting emissions under 348 nm excitation. Sharp and strong emission bands at 554 nm (green), 664 nm (red), 760 nm (near-infrared), and 2046 nm (infrared) broadband emission are observed as a result of the 4f-4f transition in Ho3+ ions. The intensity of the 554 nm emission band doubled on increasing the Ho concentration from 9.791 wt.% to 18.068 wt.% and no concentration quenching was observed. The CIE chromaticity color evaluation shows that Ho-SiAlON ceramics may be a potential candidate for green-emitting phosphors.

Author Contributions

Conceptualization: Y.K.K.; Data curation, Y.K.K. and B.C.; Formal analysis, Y.K.K. and B.C.; Funding acquisition, Y.K.K. and T.-H.K.; Investigation, Y.K.K., B.C. and D.R.D.; Methodology, Y.K.K.; Project administration, T.-H.K.; Resources, S.W.L. and T.-H.K.; Supervision, Y.K.K.; Validation, Y.K.K. and D.R.D.; Writing—original draft, Y.K.K.; Writing—review and editing, B.C., D.R.D., S.W.L. and T.-H.K. All authors have read and agreed to the published version of the manuscript.

Funding

This research was supported by the Brain Pool Program through the National Research Foundation of Korea (NRF) funded by the Ministry of Science and ICT (Grant No. 2020H1D3A1A02081359), Basic Science Research Program through the National Research Foundation of Korea (NRF) funded by the Ministry of Education (Grant No. 2021R1I1A3059543), and National Research Foundation of Korea (NRF) funded by the Ministry of Science and ICT (Grant No. 2021R1F1A1058276).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Joshi, B.; Gyawali, G.; Wang, H.; Sekino, T.; Lee, S.W. Thermal and mechanical properties of hot pressed translucent Y2O3 doped Mg–α/β-Sialon ceramics. J. Alloys Compd. 2013, 557, 112–119. [Google Scholar] [CrossRef]
  2. Ekström, T.; Persson, J. Hot Hardness Behavior of Yttrium Sialon Ceramics. J. Am. Ceram. Soc. 1990, 73, 2834–2838. [Google Scholar] [CrossRef]
  3. Hampshire, S.; Park, H.K.; Thompson, D.P.; Jack, K.H. α′-Sialon ceramics. Nature 1978, 274, 880–882. [Google Scholar] [CrossRef]
  4. Greskovich, C.; Rosolowski, J.H. Sintering of Covalent Solids. J. Am. Ceram. Soc. 1976, 59, 525. [Google Scholar] [CrossRef]
  5. Cao, G.Z.; Metselaar, R. α’-Sialon ceramics: A review. Chem. Mater. 1991, 3, 242–252. [Google Scholar] [CrossRef]
  6. Xie, R.-J.; Hirosaki, N.; Mitomo, M.; Yamamoto, Y.; Suehiro, T.; Ohashi, N. Photoluminescence of Cerium-Doped α-SiAlON Materials. J. Am. Ceram. Soc. 2004, 87, 1368–1370. [Google Scholar] [CrossRef]
  7. Suehiro, T.; Hirosaki, N.; Xie, R.-J.; Mitomo, M. Powder Synthesis of Ca-α’-SiAlON as a Host Material for Phosphors. Chem. Mater. 2004, 17, 308–314. [Google Scholar] [CrossRef]
  8. Xie, R.-J.; Hirosaki, N. Silicon-based oxynitride and nitride phosphors for white LEDs—A review. Sci. Technol. Adv. Mater. 2007, 8, 588–600. [Google Scholar] [CrossRef]
  9. Joshi, B.; Hoon, J.S.; Kshetri, Y.K.; Gyawali, G.; Lee, S.W. Transparent Sialon phosphor ceramic plates for white light emitting diodes applications. Ceram. Int. 2018, 44, 23116–23124. [Google Scholar] [CrossRef]
  10. Tang, W.; Guo, Q.; Su, K.; Liu, H.; Zhang, Y.; Mei, L.; Liao, L. Structure and Photoluminescence Properties of Dy3+ Doped Phosphor with Whitlockite Structure. Materials 2022, 15, 2177. [Google Scholar] [CrossRef]
  11. Setlur, A.; Porob, D.; Happek, U.; Brik, M. Concentration quenching in Ce3+-doped LED phosphors. J. Lumin. 2013, 133, 66–68. [Google Scholar] [CrossRef]
  12. Garcia-Santamaria, F.; Murphy, J.E.; Setlur, A.A.; Sista, S.P. Concentration Quenching in K2SiF6:Mn4+ Phosphors. ECS J. Solid State Sci. Technol. 2017, 7, R3030–R3033. [Google Scholar] [CrossRef]
  13. Wang, Z.; Meijerink, A. Concentration Quenching in Upconversion Nanocrystals. J. Phys. Chem. C 2018, 122, 26298–26306. [Google Scholar] [CrossRef]
  14. Yang, N.; Li, J.; Zhang, Z.; Wen, D.; Liang, Q.; Zhou, J.; Yan, J.; Shi, J. Delayed Concentration Quenching of Luminescence Caused by Eu3+-Induced Phase Transition in LaSc3(BO3)4. Chem. Mater. 2020, 32, 6958–6967. [Google Scholar] [CrossRef]
  15. Atabaev, T.S.; Vu, H.-H.T.; Kim, Y.-D.; Lee, J.-H.; Kim, H.-K.; Hwang, Y.-H. Synthesis and luminescence properties of Ho3+ doped Y2O3 submicron particles. J. Phys. Chem. Solids 2012, 73, 176–181. [Google Scholar] [CrossRef]
  16. Dexter, D.L.; Schulman, J.H. Theory of Concentration Quenching in Inorganic Phosphors. J. Chem. Phys. 1954, 22, 1063–1070. [Google Scholar] [CrossRef]
  17. Zhou, J.; Liu, Q.; Feng, W.; Sun, Y.; Li, F. Upconversion Luminescent Materials: Advances and Applications. Chem. Rev. 2014, 115, 395–465. [Google Scholar] [CrossRef]
  18. Wisser, M.D.; Chea, M.; Lin, Y.; Wu, D.M.; Mao, W.L.; Salleo, A.; Dionne, J.A. Strain-Induced Modification of Optical Selection Rules in Lanthanide-Based Upconverting Nanoparticles. Nano Lett. 2015, 15, 1891–1897. [Google Scholar] [CrossRef]
  19. Runowski, M.; Shyichuk, A.; Tymiński, A.; Grzyb, T.; Lavín, V.; Lis, S. Multifunctional Optical Sensors for Nanomanometry and Nanothermometry: High-Pressure and High-Temperature Upconversion Luminescence of Lanthanide-Doped Phosphates—LaPO4/YPO4:Yb3+–Tm3+. ACS Appl. Mater. Interfaces 2018, 10, 17269–17279. [Google Scholar] [CrossRef]
  20. Kshetri, Y.K.; Joshi, B.; Diaz-Torres, L.A.; Lee, S.W. Efficient Near Infrared to Visible and Near-Infrared Upconversion Emissions in Transparent (Tm3+, Er3+)-α-Sialon Ceramics. J. Am. Ceram. Soc. 2016, 100, 224–234. [Google Scholar] [CrossRef]
  21. Kshetri, Y.K.; Kamiyama, T.; Torii, S.; Jeong, S.H.; Kim, T.-H.; Choi, H.; Zhou, J.; Feng, Y.P.; Lee, S.W. Electronic structure, thermodynamic stability and high-temperature sensing properties of Er-α-SiAlON ceramics. Sci. Rep. 2020, 10, 4952. [Google Scholar] [CrossRef] [PubMed]
  22. Sun, W.-Y.; Tien, T.-Y.; Yen, T.-S. Solubility Limits of α’-SiAION Solid Solutions in the System Si,Al,Y/N,O. J. Am. Ceram. Soc. 1991, 74, 2547–2550. [Google Scholar] [CrossRef]
  23. Shen, Z.; Nygren, M. On the extension of the α-sialon phase area in yttrium and rare-earth doped systems. J. Eur. Ceram. Soc. 1997, 17, 1639–1645. [Google Scholar] [CrossRef]
  24. Grins, J.; Esmaeilzadeh, S.; Svensson, G.; Shen, Z. high-resolution electron microscopy of aSr-containing sialon polytypoid phase. J. Eur. Ceram. Soc. 1999, 19, 2723–2730. [Google Scholar] [CrossRef]
  25. Bernhardt, G.; Krassikoff, J.; Sturtevant, B.; Lad, R. Properties of amorphous SiAlON thin films grown by RF magnetron co-sputtering. Surf. Coat. Technol. 2014, 258, 1191–1195. [Google Scholar] [CrossRef]
  26. Mohamedkhair, A.; Hakeem, A.; Drmosh, Q.; Mohammed, A.; Baig, M.; Ul-Hamid, A.; Gondal, M.; Yamani, Z. Fabrication and Characterization of Transparent and Scratch-Proof Yttrium/Sialon Thin Films. Nanomaterials 2020, 10, 2283. [Google Scholar] [CrossRef]
  27. Michalik, D.; Pawlik, T.; Kukliński, B.; Lazarowska, A.; Leśniewski, T.; Barzowska, J.; Mahlik, S.; Grinberg, M.; Adamczyk, B.; Pławecki, M.; et al. Dopant Concentration Induced Optical Changes in Ca,Eu-α-Sialon. Crystals 2017, 7, 342. [Google Scholar] [CrossRef]
  28. Hagio, T.; Takase, A.; Umebayashi, S. X-ray photoelectron spectroscopic studies of β-sialons. J. Mater. Sci. Lett. 1992, 11, 878–880. [Google Scholar] [CrossRef]
  29. Kshetri, K.Y.; Chaudhary, B.; Dhakal, D.R.; Murali, G.; Pachhai, S.; Lee, S.W.; Kim, H.; Kim, T. Anomalous Upconversion Behavior and High-Temperature Spectral Properties of Yb /Ho-SiAlON Ceramics. Ceram. Int. 2022, in press. [Google Scholar] [CrossRef]
  30. Brow, R.K.; Pantano, C.G. Thermochemical Nitridation of Microporous Silica Films in Ammonia. J. Am. Ceram. Soc. 1987, 70, 9–14. [Google Scholar] [CrossRef]
  31. Donley, M.S.; Baer, D.R.; Stoebe, T.G. Nitrogen 1s charge referencing for Si3N4 and related compounds. Surf. Interface Anal. 1988, 11, 335–340. [Google Scholar] [CrossRef]
  32. Moulder, J.F.; Stickle, W.F.; Sobol, P.E.; Bomben, K.D. Handbook of X Ray Photoelectron Spectroscopy; Perkin-Elmer Corporation, Physical Electronics Division: Eden Prairie, MN, USA, 1992; ISBN 978-0964812413. [Google Scholar]
  33. Min, J.-W.; Mitomo, M. Preparation of Y-α-sialon with glassy or crystalline phases at grain boundaries. Ceram. Int. 1995, 21, 427–432. [Google Scholar] [CrossRef]
  34. Tsai, M.-H.; Wang, H.-Y.; Lu, H.-T.; Tseng, I.-H.; Lu, H.-H.; Huang, S.-L.; Yeh, J.-M. Properties of polyimide/Al2O3 and Si3N4 deposited thin films. Thin Solid Films 2011, 519, 4969–4973. [Google Scholar] [CrossRef]
  35. Pélisson-Schecker, A.; Hug, H.J.; Patscheider, J. Charge referencing issues in XPS of insulators as evidenced in the case of Al-Si-N thin films. Surf. Interface Anal. 2011, 44, 29–36. [Google Scholar] [CrossRef]
  36. Joo, M.H.; Park, S.J.; Hong, S.-M.; Rhee, C.K.; Kim, D.; Sohn, Y. Electrodeposition and Characterization of Lanthanide Elements on Carbon Sheets. Coatings 2021, 11, 100. [Google Scholar] [CrossRef]
  37. Carnall, W.T.; Fields, P.R.; Rajnak, K. Electronic Energy Levels in the Trivalent Lanthanide Aquo Ions. I. Pr3+, Nd3+, Pm3+, Sm3+, Dy3+, Ho3+, Er3+, and Tm3+. J. Chem. Phys. 1968, 49, 4424–4442. [Google Scholar] [CrossRef]
  38. Shen, Z.; Nygren, M.; Hålenius, U. Absorption spectra of rare-earth-doped α-sialon ceramics. J. Mater. Sci. Lett. 1997, 16, 263–266. [Google Scholar] [CrossRef]
  39. Mondragon, M.; Garcia, J.; Sibley, W.; Hunt, C. Optical absorption and emission from Ho3+ ions in KCaF3 crystals. J. Solid State Chem. 1988, 76, 368–374. [Google Scholar] [CrossRef]
  40. Min, X.; Fang, M.; Huang, Z.; Liu, H.; Liu, Y.; Tang, C.; Wu, X. A novel green phosphor LaMgAl11O19:Ho3+ for near-UV/blue light-pumped white light-emitting diodes. Chem. Phys. Lett. 2015, 618, 182–185. [Google Scholar] [CrossRef]
  41. Singh, V.; Dabre, K.; Dhoble, S.; Lakshminarayana, G. Green emitting holmium (Ho) doped yttrium oxide (Y2O3) phosphor for solid state lighting. Optik 2020, 206, 164339. [Google Scholar] [CrossRef]
  42. Kshetri, Y.K.; Hoon, J.S.; Kim, T.-H.; Sekino, T.; Lee, S.W. Yb3+, Er3+ and Tm3+ doped α-Sialon as upconversion phosphor. J. Lumin. 2018, 204, 485–492. [Google Scholar] [CrossRef]
  43. Wan, Z.; Li, W.; Mei, B.; Liu, Z.; Yang, Y. Fabrication and spectral properties of Ho-doped calcium fluoride transparent ceramics. J. Lumin. 2020, 223, 117188. [Google Scholar] [CrossRef]
  44. Rani, P.R.; Venkateswarlu, M.; Swapna, K.; Mahamuda, S.; Prasad, M.S.; Rao, A. Spectroscopic and luminescence properties of Ho3+ ions doped Barium Lead Alumino Fluoro Borate glasses for green laser applications. Solid State Sci. 2020, 102, 106175. [Google Scholar] [CrossRef]
  45. Bordj, S.; Satha, H.; Barros, A.; Zambon, D.; Jouart, J.-P.; Diaf, M.; Mahiou, R. Spectroscopic characterization by up conversion of Ho3+/Yb3+ codoped CdF2 single crystal. Opt. Mater. 2021, 118, 111249. [Google Scholar] [CrossRef]
  46. Kaczkan, M.; Malinowski, M. Optical Transitions and Excited State Absorption Cross Sections of SrLaGaO4 Doped with Ho3+ Ions. Materials 2021, 14, 3831. [Google Scholar] [CrossRef]
Figure 1. (a) XRD spectra of different Ho-SiAlON ceramic samples of composition presented in Table 1. The line spectrum on the bottom axis is the XRD spectra of the Ca-α-SiAlON of JCPDS 33-261. (b) Enlarged view of the (210) XRD peak.
Figure 1. (a) XRD spectra of different Ho-SiAlON ceramic samples of composition presented in Table 1. The line spectrum on the bottom axis is the XRD spectra of the Ca-α-SiAlON of JCPDS 33-261. (b) Enlarged view of the (210) XRD peak.
Materials 15 06715 g001
Figure 2. SEM images of the fracture surface of the Ho-SiAlON ceramics (a) H10, (b) H11, (c) H15, and (d) H20. The scale bar is 5 µm in all the images.
Figure 2. SEM images of the fracture surface of the Ho-SiAlON ceramics (a) H10, (b) H11, (c) H15, and (d) H20. The scale bar is 5 µm in all the images.
Materials 15 06715 g002
Figure 3. TEM micrograph of H15 sample (left) and quantitative EDS spectra (right) of the points indicated in the TEM micrograph on the left. Points 1, 3, 4, and 6 correspond to inside the grains, and points 2 and 5 correspond to the triple junctions, respectively.
Figure 3. TEM micrograph of H15 sample (left) and quantitative EDS spectra (right) of the points indicated in the TEM micrograph on the left. Points 1, 3, 4, and 6 correspond to inside the grains, and points 2 and 5 correspond to the triple junctions, respectively.
Materials 15 06715 g003
Figure 4. Elemental mapping diagram of H15 sample. (a) HAADF image. Images (bf) correspond to the mapping spectra of Si, Al, O, N and Ho, respectively, of the region indicated in (a). (g) Superposition of the elemental spectra of (bf).
Figure 4. Elemental mapping diagram of H15 sample. (a) HAADF image. Images (bf) correspond to the mapping spectra of Si, Al, O, N and Ho, respectively, of the region indicated in (a). (g) Superposition of the elemental spectra of (bf).
Materials 15 06715 g004
Figure 5. (a) XPS survey spectra of H15 sample. (bf) are the high-resolution XPS spectra of constituent elements in the H15 sample. Dots represent the observed data points and continuous lines represent the fitted graph.
Figure 5. (a) XPS survey spectra of H15 sample. (bf) are the high-resolution XPS spectra of constituent elements in the H15 sample. Dots represent the observed data points and continuous lines represent the fitted graph.
Materials 15 06715 g005
Figure 6. Absorption spectra of Ho-SiAlON samples.
Figure 6. Absorption spectra of Ho-SiAlON samples.
Materials 15 06715 g006
Figure 7. Downshifting spectra of Ho-SiAlON samples under 348 nm (a) CW excitation and (b) pulsed excitation.
Figure 7. Downshifting spectra of Ho-SiAlON samples under 348 nm (a) CW excitation and (b) pulsed excitation.
Materials 15 06715 g007
Figure 8. CIE 1931 chromaticity diagram of the Ho-SiAlON ceramics samples (a) upon 348 nm CW excitation (b) upon 348 nm pulsed excitation. Corresponding CIE coordinates are listed in Table 3. The emission modulation under CW and pulsed excitations result in different emission colors.
Figure 8. CIE 1931 chromaticity diagram of the Ho-SiAlON ceramics samples (a) upon 348 nm CW excitation (b) upon 348 nm pulsed excitation. Corresponding CIE coordinates are listed in Table 3. The emission modulation under CW and pulsed excitations result in different emission colors.
Materials 15 06715 g008
Figure 9. Downshifting spectra in the infrared range of Ho-SiAlON samples under 348 nm CW excitation.
Figure 9. Downshifting spectra in the infrared range of Ho-SiAlON samples under 348 nm CW excitation.
Materials 15 06715 g009
Figure 10. Energy level diagram of Ho3+ ion and involved downshifting photoluminescence mechanism under 348 nm excitation. Solid upward arrow represents absorption transition while downward arrows represent emission transitions. Dotted downward arrows represent non-radiative relaxation.
Figure 10. Energy level diagram of Ho3+ ion and involved downshifting photoluminescence mechanism under 348 nm excitation. Solid upward arrow represents absorption transition while downward arrows represent emission transitions. Dotted downward arrows represent non-radiative relaxation.
Materials 15 06715 g010
Table 1. Starting m and n parameters, precursor compositions, and sample codes of the synthesized SiAlON samples.
Table 1. Starting m and n parameters, precursor compositions, and sample codes of the synthesized SiAlON samples.
SamplemnSi3N4 (wt. %)AlN (wt. %)Al2O3 (wt. %)Ho2O3 (wt. %)
H101.01.072.69814.8692.6429.791
H111.11.170.79215.6192.88810.701
H151.51.066.28818.3491.26814.096
H202.01.060.37221.560018.068
Table 2. List of binding energy, full width at half maxima (FWHM), area, and proposed chemical-bond assignments. Integer identifiers are assigned with the lowest binding energy sub-peak as #1 to the sub-peaks within a given elemental peak. The unit of binding energy, FWHM, and the area are electron volts (eV).
Table 2. List of binding energy, full width at half maxima (FWHM), area, and proposed chemical-bond assignments. Integer identifiers are assigned with the lowest binding energy sub-peak as #1 to the sub-peaks within a given elemental peak. The unit of binding energy, FWHM, and the area are electron volts (eV).
ElementPeakBinding Energy (eV)FWHM (eV)Area (eV)Assignment
SiSi2p #1101.461.6047,556.38Si−N (Si3N4)
Si2p #2102.331.8719,194.46Si−O (Glassy grain boundary phase)
AlAl2p #173.811.7810,382.01Al−N (AlN rich phase and grain boundaries)
Al2p #274.391.982372.03Al−O (Al(Si)–O(N) substitution in Si3N4)
OO1s #1531.932.31114,724.90O−Al (Al(Si)–O(N) substitution in Si3N4)
O1s #2532.391.2814,043.52O−Si (Glassy grain boundary phase)
NN1s #1396.201.7310,604.92N−Al (AlN rich phase and grain boundaries)
N1s #2397.211.64117,564.10N−Si (Si3N4)
N1s #3398.862.9211,423.29N−Ho
HoHo4d #1161.321.702843.224d5/2
Ho4d #2163.401.672309.504d3/2
Table 3. CIE coordinates of Ho-SiAlON ceramic samples under 348 CW and pulsed excitations.
Table 3. CIE coordinates of Ho-SiAlON ceramic samples under 348 CW and pulsed excitations.
Excitation Wavelength, ModeCIE DiagramSample Name (Point on the CIE Diagram)CIE Coordinates
XYI (lum)
348 nm; CWFigure 6aH20 (1)0.296630.555514.724 × 104
H15 (2)0.302850.572893.338 × 104
H11 (3)0.263770.476923.928 × 104
H10 (4)0.313480.584762.533 × 104
348 nm; PulsedFigure 6bH20 (1)0.346520.627401.124 × 103
H15 (2)0.345710.632558.863 × 102
H11 (3)0.332640.630211.094 × 103
H10 (4)0.351480.628607.299 × 102
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Kshetri, Y.K.; Chaudhary, B.; Dhakal, D.R.; Lee, S.W.; Kim, T.-H. Ho-SiAlON Ceramics as Green Phosphors under Ultra-Violet Excitations. Materials 2022, 15, 6715. https://doi.org/10.3390/ma15196715

AMA Style

Kshetri YK, Chaudhary B, Dhakal DR, Lee SW, Kim T-H. Ho-SiAlON Ceramics as Green Phosphors under Ultra-Violet Excitations. Materials. 2022; 15(19):6715. https://doi.org/10.3390/ma15196715

Chicago/Turabian Style

Kshetri, Yuwaraj K., Bina Chaudhary, Dhani Ram Dhakal, Soo Wohn Lee, and Tae-Ho Kim. 2022. "Ho-SiAlON Ceramics as Green Phosphors under Ultra-Violet Excitations" Materials 15, no. 19: 6715. https://doi.org/10.3390/ma15196715

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