Next Article in Journal
Preparation, Characterization, and Chemically Modified Date Palm Fiber Waste Biomass for Enhanced Phenol Removal from an Aqueous Environment
Next Article in Special Issue
Fabrication and Luminescent Properties of Highly Transparent Er:Y2O3 Ceramics by Hot Pressing Sintering
Previous Article in Journal
Heat Treatment Optimization for a High Strength Al–Mn–Sc Alloy Fabricated by Selective Laser Melting
Previous Article in Special Issue
Structural and Scintillation Properties of Ce3+:Gd3Al3Ga2O12 Translucent Ceramics Prepared by One-Step Sintering
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Study on Local-Structure Symmetrization of K2XF6 Crystals Doped with Mn4+ Ions by First-Principles Calculations

1
Postgraduate Program of Science Education, Universitas PGRI Semarang, Semarang 50232, Indonesia
2
Faculty of Mathematics, Natural Sciences and Information Technology Education, Universitas PGRI Semarang, Semarang 50232, Indonesia
3
Faculty of Engineering and Informatics, Universitas PGRI Semarang, Semarang 50232, Indonesia
4
Faculty of Pharmacy, Universitas Setia Budi, Surakarta 57127, Indonesia
5
Department of Physics, Universitas Kristen Satya Wacana, Salatiga 50711, Indonesia
6
Department of Computer Application, Galgotias University, Greater Noida 203201, India
7
Independent Researcher, 925 Dalney Street NW, Atlanta, GA 30318, USA
8
School of Science and Technology, Kwansei Gakuin University, Sanda 669-1337, Japan
9
Department of Theoretical Physics, Jan Dlugosz University, 42-200 Czestochowa, Poland
10
Centre of Excellence for Photoconversion, Vinča Institute of Nuclear Sciences-National Institute of the Republic of Serbia, University of Belgrade, 11351 Belgrade, Serbia
11
School of Optoelectronic Engineering and CQUPT-BUL Innovation Institute, Chongqing University of Posts and Telecommunications, Chongqing 400065, China
12
Academy of Romanian Scientists, Ilfov Str. No. 3, 010071 Bucharest, Romania
13
Institute of Physics, University of Tartu, W. Ostwald Str. 1, 50411 Tartu, Estonia
14
Institute of Solid State Physics, University of Latvia, Kengaraga 8, LV-1063 Riga, Latvia
*
Author to whom correspondence should be addressed.
Materials 2023, 16(11), 4046; https://doi.org/10.3390/ma16114046
Submission received: 30 April 2023 / Revised: 19 May 2023 / Accepted: 26 May 2023 / Published: 29 May 2023
(This article belongs to the Special Issue Glasses and Ceramics for Luminescence Applications)

Abstract

:
The crystals of Mn4+-activated fluorides, such as those of the hexafluorometallate family, are widely known for their luminescence properties. The most commonly reported red phosphors are A2XF6: Mn4+ and BXF6: Mn4+ fluorides, where A represents alkali metal ions such as Li, Na, K, Rb, Cs; X=Ti, Si, Ge, Zr, Sn, B = Ba and Zn; and X = Si, Ge, Zr, Sn, and Ti. Their performance is heavily influenced by the local structure around dopant ions. Many well-known research organizations have focused their attention on this area in recent years. However, there has been no report on the effect of local structural symmetrization on the luminescence properties of red phosphors. The purpose of this research was to investigate the effect of local structural symmetrization on the polytypes of K2XF6 crystals, namely Oh-K2MnF6, C3v-K2MnF6, Oh-K2SiF6, C3v-K2SiF6, D3d-K2GeF6, and C3v-K2GeF6. These crystal formations yielded seven-atom model clusters. Discrete Variational Xα (DV-Xα) and Discrete Variational Multi Electron (DVME) were the first principles methods used to compute the Molecular orbital energies, multiplet energy levels, and Coulomb integrals of these compounds. The multiplet energies of Mn4+ doped K2XF6 crystals were qualitatively reproduced by taking lattice relaxation, Configuration Dependent Correction (CDC), and Correlation Correction (CC) into account. The 4A2g4T2g (4F) and 4A2g4T1g (4F) energies increased when the Mn-F bond length decreased, but the 2Eg4A2g energy decreased. Because of the low symmetry, the magnitude of the Coulomb integral became smaller. As a result, the decreasing trend in the R-line energy could be attributed to a decreased electron–electron repulsion.

1. Introduction

Incandescent and fluorescent lighting sources have been rapidly replaced by White Light Emitting Diodes (WLEDs) in homes, offices, and public areas. They are made of a blue LED chip with a yellow phosphor. WLED is the most energy-efficient conversion source compared to previously existing lighting sources, yet it creates pseudo-white light due to a lack of red emissions. Rather than employing all basic colors of LED chips, mixing blue LED chips with yellow and red phosphors is simpler and less expensive. The blue LED chips are typically made of InGaN [1], whereas the yellow phosphor components are composed of Y3Al5O12: Ce3+ [2]. The high-performance red phosphors are Eu2+ doped nitrides [3,4,5,6,7,8,9]. Unfortunately, red phosphors are expensive due to scarcity and challenging synthesis conditions, such as extreme temperatures and nitrogen pressure. Finding novel red phosphor materials that are appropriate for WLED is currently challenging. Significant performance factors for white light that are used in general lighting include high Quantum Efficiency (QE > 70%), resistance to thermal quenching (preferably > 80% of the luminescence intensity should be sustained at 450 K), and strong color quality, which includes a low Correlated Color Temperature (CCT) of 3000 K and a high Color Rendering Index (CRI > 70).
The most commonly reported red phosphors are fluoride-based, such as A2XF6: Mn4+ and BXF6: Mn4+, where A represents alkali metal ions, such as Li, Na, K, Rb, Cs; X = Ti, Si, Ge, Zr, Sn), B = Ba and Zn and X = Si, Ge, Zr, Sn, and Ti. K2SiF6: Mn4+, KNa2SiF6: Mn4+, and K2TiF6: Mn4+, in particular, have shown good potential for WLED as a red phosphor under blue LED chip stimulation. The first red Mn4+-doped fluoride phosphor, K2SiF6: Mn4+, was published in 1973 [10]. K2SiF6 is one of the most promising hexafluoride hosts, with a slightly higher Luminous Efficacy of Radiation (LER) upon Mn4+ doping than K2TiF6 and a 30% higher External Quantum Efficiency (EQE) than KnaSiF6: Mn4+ [11]. Mn4+, when doped in K2SiF6 or K2TiF6 as a red phosphor, yields WLEDs with warm-white CCTs ~3000 K and good CRIs ~90, as demonstrated by Setlur et al. [12]. The d–d transitions in Mn4+ cause the particular red emission line detected in K2SiF6: Mn4+ to be approximately 630 nm (15,873 cm−1 or 1.97 eV) [13]. Nevertheless, the chemical and thermal stability problems and safety hazards of K2SiF6 and K2TiF6 doped with Mn4+ have been reported.
The aforementioned red-phosphor performance is highly dependent on local structure. Numerous research teams have concentrated on the modification and enhancement of phosphor luminescence properties through the alteration of the local crystal structure. The “Cation-Size-Mismatch effect”, “Neighboring-Cation Substitution effect”, and “Nanosegregation and Neighbor-Cation Control effect”, among other new luminescence mechanisms, were reported by Liu’s group in Ce3+ and Eu2+-doped (oxy)nitrides based on the variation in the local crystal structure [14,15,16,17]. Ram’s team also showed that slight modifications to the local structure of phosphor systems such as La3xCexSi6N11, SrxBa2xSiO4: Eu2+, etc., could lead to appreciable gains in luminescence performance [18,19]. Cheetham’s team discovered that local crystal structural deformation accounted for a significant spectrum change from blue to yellow light from Ca2SiO4: Ce [20].
The Ligand Field Theory (LFT) has been frequently used to successfully evaluate the multiplet energy levels and optical spectra of Transition Metal (TM) ions in crystals [21]. However, it is an empirical method in which the measured spectrum is used to determine the Racah parameters and crystal field splitting. Watanabe and Kamimura produced the first non-empirical forecast in the late 1980s [22,23] using a combination of the local density approximation (LDA) and LFT. On the other hand, a number of teams, including Daul et al. [24], Wissing et al. [25], and Oliveira et al. [26,27], have also performed first-principle calculations based on the Density Functional Theory (DFT). However, obtaining the many-electron wave functions proved unfeasible. During the previous ten years, Ogasawara’s team created the Discrete Variational Multi-Electron (DVME) approach [28]: a non-empirical first-principles many-electron calculation technique. It uses both a Configuration Interaction (CI) computation and DFT. DVME consists of two phases. To begin, one-electron Molecular Orbital (MO) calculations are performed using the Discrete Variational Xα (DV-Xα) method. The CI method is then used to perform many-electron computations, which is the main stage of the DVME approach. It has been shown that DVME is a powerful tool for estimating absorption spectra, energy levels, transition energies, etc., without the use of any empirical parameters.
Up until recently, there has been no study on the influence of local structure on symmetry (switching from a high-symmetry to a low-symmetry configuration) or on the luminous qualities of red phosphors. Therefore, the goal of this research was to investigate the effect of local structural symmetrization on the polytypes of K2XF6 crystals, namely Oh−-K2MnF6, C3v-K2MnF6, Oh-K2SiF6, C3v-K2SiF6, D3d-K2GeF6, and C3v-K2GeF6. The DVME method was used to calculate their multiplet energy levels.

2. Materials and Methods

Polytypes of various K2XF6 crystals were used to create seven-atom model clusters. The cubic K2MnF6 ICSD #47213 had a = 8.221 lattice parameter, a space group F m 3 ¯ m , and Oh symmetry [29]. The lattice parameters of hexagonal K2MnF6 ICSD #60417 were a = 5.719 Å and c = 9.330 Å, with space group P 63 m c and C3v Symmetry [30]. The cubic K2SiF6 ICSD #2940 had a= 8.134 lattice parameters, a space group F m 3 ¯ m , and Oh symmetry [31]. The lattice parameters of hexagonal K2SiF6 ICSD #158483 were a = 5.6461 Å and c = 9.2322 Å, with the space group P 63 m c and C3v Symmetry [32]. The lattice parameters of rhombohedral K2GeF6 ICSD #24026 were a = 5.63 and c = 4.66, with the space group P 3 ¯ m 1 and D3d Symmetry [33]. The lattice parameters of hexagonal K2GeF6 ICSD #30310 were a = 5.71 Å and c = 9.27 Å, with the space group P 63 m c and C3v Symmetry [34]. The computations were performed using Oh, D3d, and C3v symmetry for clusters built from K2XF6 (X = Mn, Si, or Ge) and crystals with cubic, rhombohedral, and hexagonal structures, respectively. Figure 1a–c depicts the various types of crystal structures of the materials under consideration, including namely cubic, rhombohedral, and hexagonal structures. Figure 1d–f were model clusters made up of seven atoms, one X4+ ion surrounded by 6 F. Here, we adopted the results of the Mn K-edge Extended X-ray Absorption Fine Structure (EXAFS) measurement of K2XF6 (X = Si, or Ge): Mn4+ [35]. The Mn-F bond lengths for K2SiF6: Mn4+ and K2GeF6: Mn4+ were 1.807 and 1.810 Å, respectively. The one-electron calculations utilizing the DV-Xα method were then carried out [36,37,38]. The DVME approach was used to account for the many-electron effects [28]. The energy corrections such as Configuration Dependent Correction (CDC) and Correlation Correction (CC) were also considered. Racah parameters were used to calculate the Coulomb integrals as well. These methods’ specific steps are described in Reference [35].

3. Results

3.1. Bond Lengths

The Mn-F bond lengths of Oh-K2MnF6, C3v-K2MnF6, Oh-K2SiF6, C3v-K2SiF6, D3d-K2GeF6, and C3v-K2GeF6. are shown in Table 1. All six bond lengths are represented by letters d1, d2, d3, d4, d5, and d6, respectively. When the lattice relaxation effect was not used, the lengths of the Mn-F bonds dropped from Oh-K2MnF6 to C3v-K2MnF6. This was similar to the trend for Oh-K2SiF6: Mn4+ to C3v-K2SiF6: Mn4+. On the other hand, the trend for D3d-K2GeF6: Mn4+ to C3v-K2GeF6: Mn4+ was reversed. When the lattice relaxation effect was used, however, the Mn-F bond lengths decreased in all situations.

3.2. Molecular Orbital Energies

Figure 2 depicts the molecular orbital energies of Oh-K2MnF6, C3v-K2MnF6, Oh-K2SiF6, C3v-K2SiF6, D3d-K2GeF6, and C3v-K2GeF6. The Valence Band (VB) is represented by black solid lines. The Conduction Band (CB) is shown by the black dashed lines. The impurity levels are represented as t2g and eg, with solid red and dashed blue lines, respectively. The tops of the VBs were set to zero. For Oh-K2MnF6 and C3v-K2MnF6, the crystal field splitting (10Dq, defined as the differential energy between t2g and eg levels) was estimated to be 1.79 and 2.68 eV, respectively. Without accounting for the lattice relaxation effect, the 10Dq of Oh-K2SiF6 and C3v-K2SiF6 were estimated to be 3.52 and 3.44 eV, respectively. They fell to 2.63 and 2.53 eV when the lattice relation effect was taken into account. In the case of D3d-K2GeF6 and C3v-K2GeF6, the 10Dq was determined to be 2.76 and 2.72 eV, respectively. After accounting for the lattice relaxation effect, they fell to 2.52 and 2.61 eV, respectively.

3.3. Multiplet Energy Levels

Since the d–d transitions of K2XF6: Mn4+ was prohibited by the parity selection rule, the transition probabilities could not be determined. As a result, this report is restricted to energy levels. We estimated the doublet states 2E,.2T2, and 2T1, as well as the quartet states 4T2 and 4T1a. The absorption transitions start from the ground 4A2 state to 4T2 and 4T1a states which often appeared as wide bands and were referred to as the U- and Y-band, respectively. On the other hand, the emission transition started from the 2E state to the ground 4A2 state, which generally appeared as a sharp line and was referred to as R-line.
The pure K2MnF6 and K2SiF6: Mn4+ computed multiplet energy diagrams with Oh and C3v symmetry are shown in Figure 3. A few adjustments, including CDC, CC, and lattice relaxation, were also assessed. Figure 3 demonstrates that quite often, the doublet states decreased when reduced symmetry was employed. Furthermore, CDC-CC correction had a smaller impact on Oh-K2MnF6 than it did on C3v-K2MnF6, suggesting that C3v-K2MnF6 benefited more from correlation correction. On the other hand, the quartet states increased for pure K2MnF6 from Oh to C3v while they dropped for K2SiF6: Mn4+ in the same order. This was expected because the Mn-F bond length, which varied widely depending on the material, primarily affected the quartet states.
The predicted multiplet energy diagrams of K2GeF6: Mn4+ with D3d and C3v symmetry are shown in Figure 4. CDC, CC, and lattice relaxation were also evaluated, similar to Figure 3. These findings showed that the average doublet state values for the two clusters were remarkably similar. Low symmetry was also found to have an impact on multiplet splitting. While the splitting of the 4T2 state decreased, it increased for the 4T1a, 2T2, and 2T1 states.

3.4. Coulomb Integrals

The Coulomb integrals of pure K2MnF6, K2SiF6: Mn4+, and K2GeF6: Mn4+ are shown in Table 2. When low symmetry was used, the effective Coulomb integrals J eff estimated by c λ J AO almost always decreased. Although the J eff t 2 g of K2GeF6: Mn4+ without lattice relaxation was greater than that of D3d-K2GeF6: Mn4+, its tendency improved when lattice relaxation was considered. These findings suggest that reduced symmetry resulted in a smaller Coulomb integral. As a result, the decreasing trend of R-line energy could be attributed to a decreased electron–electron repulsion.

4. Discussion

LEDs are now used in a variety of commonplace applications, including display backlights for smartphones, tablets, and televisions, as well as warm-white LEDs for energy-efficient lighting. A portion of the blue light from the LED chip is converted into white light by color-converting luminescent materials. This is accomplished by using doped wide-bandgap materials, also referred to as phosphors or Colloidal Quantum Dots (QDs). The color quality of white LEDs is improved when red-emitting phosphors are added when compared to the prototype’s arrangement of a blue LED and a yellow Y3Al5O12: Ce3+. These luminous materials have an incredibly high luminescence efficiency, especially at room temperature and above, due to the involvement and stimulation of thermal phonons.
A rare-earth ion such as Eu2+ and Ce3+ or a transition metal such as Mn2+ and Mn4+ are doped into an inorganic host material to create phosphor materials. Rare-earth ions are frequently used in conventional LED phosphors, with the main component being Y3Al5O12: Ce3+ (YAG: Ce) [39]. By altering the host compound’s composition, the dopant Eu2+ can change the emission spectrum, for example, from a green emission in SrGa2S4: Eu2+ [3,4] and SrSi2O2N2: Eu2+ [5] to a red emission in Sr2Si5N8: Eu2+ [5,6], (Ca, Sr)S: Eu2+ [3,7], CaAlSiN3: Eu2+ [8] and Sr[LiAl3N4]: Eu2+ [9]. A current trend toward creating non-rare-earth element LED phosphors is the result of environmental challenges, including the scarcity of rare-earth materials.
The requirements for a phosphor to be suitable for LED applications were described by Smet et al. [39] in detail. According to the Color Quality Scale (CQS) [40] or the CRI [41], the resulting white light source had a high color rendering. This was significant for illumination. For display applications to produce a broad color spectrum or high color purity, saturated colors were necessary. The lower the filtering losses, the better the phosphors’ emission spectrum fits the color filters. Second, a phosphor must have a high LER, which is a metric for the average eye sensitivity of the spectrum, measured in lm/W) and a high Internal Quantum Efficiency (IQE), which is stable at high temperatures. Third, there needs to be significant blue light absorption, which raises the EQE. A phosphor can only be considered a serious candidate for LED applications when all four requirements are satisfied simultaneously.
The Mn4+ emission center prefers to remain in the octahedral or modified octahedral position of the host due to the large ligand-field stabilizing energy of Mn4+ in the six-fold coordination. In the initial LFT simulation, only the octahedral (Oh) crystal field was taken into account [42]. The doubly degenerate eg level had +6Dq more energy than the fivefold degenerate 3d level, and the triply degenerate t2g level had −4Dq more energy. The intensity of the crystal field, Dq, changed based on the ion-crystal combination, and as a result, the crystal field splitting was 10Dq. Through the use of the electron–electron repulsion parameters A, B, and C, also referred to as Racah parameters, the impact of covalency could also be taken into account in this situation. The crystal field strength Dq, the Racah parameters B and C, and the multiplet energy levels Ei could thus be used to explain them. The recognized Tanabe Sugano diagrams [21,43], which depict Ei/B as functions of Dq/B for a fixed value of C/B, describe the energy levels of all dN systems in an octahedral crystal field as functions of Dq. The spectral characteristics of red phosphor materials were also described using the absorption and emission spectra. There were certain doublet states, such as 2E, 2T1, etc., and some quartet states, such as 4A2, 4T2, 4T1a, 4T1b, etc., since Mn4+-doped compounds contained three electrons filling ten degenerate 3d orbitals (3d3). The energy was lowest in the ground state (4A2). The transitions from 4A2 to 4T2 (U-band) and 4T1a (Y-band) were utilized for absorption, while the transition from 2E to 4A2 (R-line) was employed for emission when used as red phosphor materials.
Previously, we studied the potential of oxide and fluoride materials for red phosphor materials in WLED by DV-Xα and DVME methods. The investigation included lattice relaxation, orbital energy, multiplet energy, absorption spectra, energy correction, pressure dependence, and emitted light utilizing CIE 1931 color space [35,44,45,46,47,48,49,50,51,52]. Because of the length of the Mn-F bond, the quartet state energies typically had a strong relationship with crystal field splitting. On the other hand, doublet state energies strongly depended on the correlation correction. The computational conditions to reproduce those optical properties depend on the material itself.
The study of low-structure symmetrization was required when understanding the properties of novel phosphor materials. According to our findings, a low structure had a substantial effect on multiplet structures, which could affect their performance as phosphor materials. Table 1 indicates that considering the lattice relaxation effect caused by the Mn4+, substitution resulted in a considerable shift in the bond lengths. The crystal field splittings were approximated using the one-electron DV-Xα approach, as shown in Figure 2. The 10Dq crystal field splitting tendency was caused by the lengths of the Mn-F bonds. Furthermore, the multiplet energies in Figure 3 and Figure 4 were determined using the many-electron DVME approach. The splitting of the corresponding multiplet energy levels was visible in lower structure symmetrization.

5. Conclusions

By taking into consideration lattice relaxation, CDC, and CC, the multiplet energies of Mn4+ doped K2XF6 crystals were qualitatively estimated. The fluoride compounds exhibited here were suitable materials to be used as red phosphors for white LEDs since the Mn4+ impurities in these hosts were emitted at approximately 620−630 nm (the proper spectral range to obtain a “warm” white light from the LEDs). In addition, they appeared to be thermally stable since they are known as commercial red phosphors (especially K2SiF6: Mn4+). We found that the Mn-F bond length dropped, yet the U- and Y-band energies increased. By contrasting various fluoride crystal symmetry polytypes, the impact of lower symmetry was explored. The Mn-F bond length decreased, while the absorption energies of 4A2g4T2g (4F) and 4A2g4T1g (4F) increased, yet the 2Eg4A2g emission energy decreased. A reduced Coulomb integral result was produced when symmetry was low. It followed that a decrease in electron–electron repulsion was the cause of the declining trend in R-line energy. The CC factor c dominated the R-line energy, while the Mn-F bond length and crystal field splitting was principally responsible for the U- and Y-band energies. We discovered that low symmetry decreased the attraction between the electrons.

Author Contributions

Conceptualization, M.N.; methodology, M.N.; software, K.O.; validation, K.O. and M.G.B.; formal analysis, D.M. and M.N.; resources, A.S.C.; data curation, M.N.; writing—original draft preparation, M.N.; writing—review and editing, B.W.; visualization, F.S.R.; supervision, M.G.B. and S.S.; project administration, S.R. and E.S.; funding acquisition, M.P. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

No new data were created.

Acknowledgments

We appreciate the support of our students who helped this research, i.e., Joko Setiawan (Magister of Science Education, Universitas PGRI Semarang), and Ammar Amjad (Department of Informatics, Universitas PGRI Semarang).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Nakamura, S. Background Story of the Invention of Efficient InGaN Blue-Light-Emitting Diodes (Nobel Lecture). Angew. Chemie Int. Ed. 2015, 54, 7770–7788. [Google Scholar] [CrossRef]
  2. Blasse, G.; Bril, A. A new phosphor for flying-spot cathode-ray tubes for color television: Yellow-emitting Y3Al5O12-Ce3+. Appl. Phys. Lett. 1967, 11, 53–55. [Google Scholar] [CrossRef]
  3. Wu, H.; Zhang, X.; Guo, C.; Xu, J.; Wu, M.; Su, Q. Three-band white light from InGaN-based blue LED chip precoated with green/red phosphors. IEEE Photonics Technol. Lett. 2005, 17, 1160–1162. [Google Scholar] [CrossRef]
  4. Joos, J.J.; Meert, K.W.; Parmentier, A.B.; Poelman, D.; Smet, P.F. Thermal quenching and luminescence lifetime of saturated green Sr1-xEuxGa2S4 phosphors. Opt. Mater. 2012, 34, 1902–1907. [Google Scholar] [CrossRef]
  5. Mueller-mach, R.; Mueller, G.; Krames, M.R.; Höppe, H.A.; Stadler, F.; Schnick, W.; Juestel, T.; Schmidt, P. Highly efficient all-nitride phosphor-converted white light emitting diode. Phys. Status Solidi 2005, 202, 1727–1732. [Google Scholar] [CrossRef]
  6. Horikawa, T.; Piao, X.Q.; Fujitani, M.; Hanzawa, H.; Machida, K. Preparation of Sr2Si5N8:Eu2+ phosphors using various novel reducing agents and their luminescent properties. IOP Conf. Ser. Mater. Sci. Eng. 2009, 1, 012024. [Google Scholar] [CrossRef]
  7. Hu, Y.; Zhuang, W.; Ye, H.; Zhang, S.; Fang, Y.; Huang, X. Preparation and luminescent properties of (Ca1-x,Srx)S:Eu2+ red-emitting phosphor for white LED. J. Lumin. 2005, 111, 139–145. [Google Scholar] [CrossRef]
  8. Uheda, K.; Hirosaki, N.; Yamamoto, Y.; Naito, A.; Nakajima, T.; Yamamoto, H. Luminescence properties of a red phosphor, CaAlSiN3:Eu2+, for white light-emitting diodes. Electrochem. Solid-State Lett. 2006, 9, H22. [Google Scholar] [CrossRef]
  9. Pust, P.; Weiler, V.; Hecht, C.; Tücks, A.; Wochnik, A.S.; Henß, A.K.; Wiechert, D.; Scheu, C.; Schmidt, P.J.; Schnick, W. Narrow-band red-emitting Sr[LiAl3N4]:Eu2+ as a next-generation LED-phosphor material. Nat. Mater. 2014, 13, 891–896. [Google Scholar] [CrossRef]
  10. Paulusz, A.G. Efficient Mn(IV) Emission in Fluorine Coordination. J. Electrochem. Soc. 1973, 120, 942. [Google Scholar] [CrossRef]
  11. Liu, R.; Nguyen, H. Narrow-band red-emitting Mn4+-doped hexafluoride phosphors: Synthesis, optoelectronic properties, and applications in white light-emitting diodes. J. Mater. Chem. C 2016, 4, 10759–10775. [Google Scholar]
  12. Setlur, A.A.; Emil, V.R.; Claire, S.H.; Her, J. -H.; Alok, M.S.; Nagaveni Karkada, M.; Satya Kishore, N.; Kumar, P.; Aesram, D.; Deshpande, A.; et al. Energy-efficient, high-color-rendering LED lamps using oxyfluoride and fluoride phosphors. Chem. Mater. 2010, 22, 4076–4082. [Google Scholar] [CrossRef]
  13. Takahashi, T.; Adachi, S. Mn4+-Activated Red Photoluminescence in K2SiF6 Phosphor. J. Electrochem. Soc. 2008, 155, E183. [Google Scholar] [CrossRef]
  14. Li, G.; Lin, C.C.; Chen, W.T.; Molokeev, M.S.; Atuchin, V.V.; Chiang, C.Y.; Zhou, W.; Wang, C.W.; Li, W.H.; Sheu, H.S.; et al. Photoluminescence tuning via cation substitution in oxonitridosilicate phosphors: DFT calculations, different site occupations, and luminescence mechanisms. Chem. Mater. 2014, 26, 2991–3001. [Google Scholar] [CrossRef]
  15. Chen, W.T.; Sheu, H.S.; Liu, R.S.; Attfield, J.P. Cation-size-mismatch tuning of photoluminescence in oxynitride phosphors. J. Am. Chem. Soc. 2012, 134, 8022–8025. [Google Scholar] [CrossRef]
  16. Wang, S.S.; Chen, W.T.; Li, Y.; Wang, J.; Sheu, H.S.; Liu, R.S. Neighboring-cation substitution tuning of photoluminescence by remote-controlled activator in phosphor lattice. J. Am. Chem. Soc. 2013, 135, 12504–12507. [Google Scholar] [CrossRef]
  17. Huang, W.Y.; Yoshimura, F.; Ueda, K.; Shimomura, Y.; Sheu, H.S.; Chan, T.S.; Greer, H.F.; Zhou, W.; Hu, S.F.; Liu, R.S.; et al. Nanosegregation and neighbor-cation control of photoluminescence in carbidonitridosilicate phosphors. Angew. Chemie Int. Ed. 2013, 52, 8102–8106. [Google Scholar] [CrossRef]
  18. George, N.C.; Denault, K.A.; Seshadri, R. Phosphors for solid-state white lighting. Annu. Rev. Mater. Res. 2013, 43, 481–501. [Google Scholar] [CrossRef]
  19. Brinkley, S.E.; Pfaff, N.; Denault, K.A.; Zhang, Z.; Hintzen, H.T.B.; Seshadri, R.; Denbaars, S.P. Robust thermal performance of Sr2Si5N8:Eu2+: An efficient red emitting phosphor for light emitting diode based white lighting. Appl. Phys. Lett. 2011, 99, 241106. [Google Scholar] [CrossRef]
  20. Kalaji, A.; Mikami, M.; Cheetham, A.K. Ce3+-Activated Γ-Ca2SiO4 and other olivine-type ABXO4 phosphors for solid-state lighting. Chem. Mater. 2014, 26, 3966–3975. [Google Scholar] [CrossRef]
  21. Tanabe, Y.; Sugano, S. On the Absorption Spectra of Complex Ions. J. Phys. Soc. Japan 1954, 9, 766–779. [Google Scholar] [CrossRef]
  22. Watanabe, S.; Kamimura, H. Multiplet Structures of Transition Metal Deep Impurities in ZnS. J. Phys. Soc. Japan 1987, 56, 1078–1091. [Google Scholar] [CrossRef]
  23. Watanabe, S.; Kamimura, H. First-principles calculations of multiplet structures of transition metal deep impurities in II-VI and III-V semiconductors. Mater. Sci. Eng. B 1989, 3, 313–324. [Google Scholar] [CrossRef]
  24. Daul, C. Density functional theory applied to the excited states of coordination compounds. Int. J. Quantum Chem. 1994, 52, 867–877. [Google Scholar] [CrossRef]
  25. Wissing, K.; Barriuso, M.T.; Aramburu, J.A.; Moreno, M. Optical excitations and coupling constants in FeO2−4 and CrO4−4 complexes in oxides: Density functional study. J. Chem. Phys. 1999, 111, 10217–10228. [Google Scholar] [CrossRef]
  26. Oliveira, M.J.T.; Medeiros, P.V.C.; Sousa, J.R.F.; Nogueira, F.; Gueorguiev, G.K. Optical and magnetic excitations of metal-encapsulating si cages: A systematic study by time-dependent density functional theory. J. Phys. Chem. C 2014, 118, 11377–11384. [Google Scholar] [CrossRef]
  27. De Oliveira, M.I.A.; Rivelino, R.; de Brito Mota, F.; Kakanakova-Georgieva, A.; Gueorguiev, G.K. Optical properties of organosilicon compounds containing sigma-electron delocalization by quasiparticle self-consistent GW calculations. Spectrochim. Acta-Part A Mol. Biomol. Spectrosc. 2021, 245, 118939. [Google Scholar] [CrossRef]
  28. Ogasawara, K.; Ishii, T.; Tanaka, I.; Adachi, H. Calculation of multiplet structures of Cr3+ and V3+ in α-Al2O3 based on a hybrid method of density-functional theory and the configuration interaction. Phys. Rev. B 2000, 61, 143–161. [Google Scholar] [CrossRef]
  29. Hoppe, R.; Hofmann, B. Neues über K2[MnF6], Rb2[MnF6] und Cs2[MnF6]. ZAAC-J. Inorg. Gen. Chem. 1977, 436, 94861268. [Google Scholar]
  30. Bukovec, P.; Hoppe, R. Zur kenntnis von hexagonalem K2[MnF6] [1]. J. Fluor. Chem. 1983, 23, 579–587. [Google Scholar] [CrossRef]
  31. Loehlin, J.H. Redetermination of the structure of potassium hexafluorosilicate, K2SiF6. Acta Crystallogr. Sect. C 1984, 40, 570. [Google Scholar] [CrossRef]
  32. Gramaccioli, C.M.; Campostrini, I. Demartinite, a new polymorph of K2SiF6 from La Fossa Crater, Vulcano, Aeolian Islands, Italy. Can. Mineral. 2007, 45, 1275–1280. [Google Scholar] [CrossRef]
  33. Hoard, J.L.; Vincent, W.B. Structures of Complex Fluorides. Potassium Hexafluogermanate and Ammonium Hexafluogermanate. J. Am. Chem. Soc. 1939, 61, 2849–2852. [Google Scholar] [CrossRef]
  34. Bode, H.; Wendt, W. Über Die Struktur von Hexafluoromanganaten(IV). ZAAC-J. Inorg. Gen. Chem. 1952, 269, 165–172. [Google Scholar]
  35. Novita, M.; Honma, T.; Hong, B.; Ohishi, A.; Ogasawara, K. Study of multiplet structures of Mn4+ activated in fluoride crystals. J. Lumin. 2016, 169, 594–600. [Google Scholar] [CrossRef]
  36. Adachi, H.; Tsukada, M.; Satoko, C. Discrete variational Xα cluster calculations. I. Application to metal clusters. J. Phys. Soc. Japan 1978, 45, 875–883. [Google Scholar] [CrossRef]
  37. Tanabe, T.; Adachi, H.; Imoto, S. Hartree-Fock-Slater Model Cluster Calculations. II. Hydrogen Chemisorption on Transition Metal Surfaces. Jpn. J. Appl. Phys. 1978, 17, 49. [Google Scholar] [CrossRef]
  38. Adachi, H.; Shiokawa, S.; Tsukada, M.; Satoko, C.; Sugano, S. Discrete variational xα cluster calculations. III. Application to transition metal complexes. J. Phys. Soc. Japan 1979, 47, 1528–1537. [Google Scholar] [CrossRef]
  39. Smet, P.F.; Parmentier, A.B.; Poelman, D. Selecting Conversion Phosphors for White Light-Emitting Diodes. J. Electrochem. Soc. 2011, 158, R37. [Google Scholar] [CrossRef]
  40. Davis, W.; Ohno, Y. Color quality scale. Opt. Eng. 2010, 49, 033602. [Google Scholar] [CrossRef]
  41. Nickerson, D. Light Sources and Color Rendering. J. Opt. Soc. Am. 1960, 50, 57–69. [Google Scholar] [CrossRef]
  42. Sugano, S.; Tanabe, Y.; Kamimura, H. Multiplets of Transition- Metal Ions in Crystals; Academic Press: New York, NY, USA, 1970. [Google Scholar]
  43. Tanabe, Y.; Sugano, S. On the Absorption Spectra of Complex Ions, III the Calculation of the Crystalline Field Strength. J. Phys. Soc. Japan 1956, 11, 864–877. [Google Scholar] [CrossRef]
  44. Novita, M.; Ogasawara, K. Comparative Study of Multiplet Structures of Mn4+ in K2SiF6, K2GeF6, and K2TiF6 Based on First-Principles Configuration–Interaction Calculations. Jpn. J. Appl. Phys. 2012, 51, 022604. [Google Scholar] [CrossRef]
  45. Novita, M.; Ogasawara, K. Comparative Study of Absorption Spectra of V2+, Cr3+, and Mn4+ in α-Al2O3 Based on First-Principles Configuration–Interaction Calculations. J. Phys. Soc. Japan 2012, 81, 104709. [Google Scholar] [CrossRef]
  46. Novita, M.; Ogasawara, K. Study on multiplet energies of V2+, Cr3+, and Mn4+ in MgO host crystal based on first-principles calculations with consideration of lattice relaxation. J. Phys. Soc. Japan 2014, 83, 124707. [Google Scholar] [CrossRef]
  47. Novita, M.; Yoshida, H.; Ogasawara, K. Investigation of ion dependence of electronic structure for 3d3 ions in Mg2TiO4based on first-principles calculations. In Proceedings of the ECS Transactions, Honolulu, HI, USA, 7–12 October 2012; Volume 50, pp. 9–17. [Google Scholar]
  48. Novita, M.; Marlina, D.; Cholifah, N.; Ogasawara, K. Study on the molecular orbital energies of ruby under pressure. Opt. Mater. 2020, 109, 110375. [Google Scholar] [CrossRef]
  49. Novita, M.; Marlina, D.; Cholifah, N.; Ogasawara, K. Enhance electron-correlation effect on the ruby multiplet energy dependence on pressure. Opt. Mater. 2020, 110, 110520. [Google Scholar] [CrossRef]
  50. Novita, M.; Farikhah, I.; Dwi Ujianti, R.M.; Marlina, D.; Walker, B.; Kiyooka, H.; Takemura, S.; Ogasawara, K. Chromaticity coordinates of ruby based on first-principles calculation. Opt. Mater. 2021, 121, 111539. [Google Scholar] [CrossRef]
  51. Novita, M.; Ujianti, R.M.D.; Nurdyansyah, F.; Supriyadi, S.; Marlina, D.; Lestari, R.A.S.; Walker, B.; Binti Mohd Razip, N.I.; Kiyooka, H.; Takemura, S.; et al. Color coordination of emerald on CIE color space based on first-principles calculations. Opt. Mater. X 2022, 16, 100184. [Google Scholar] [CrossRef]
  52. Novita, M.; Marlina, D.; Ogasawara, K.; Seok, K.J.; Soo, K.Y. Study on the optical luminescence properties of Li2Tio3: Mn4+ and Cr3+. Chem. Lett. 2021, 50, 52–56. [Google Scholar] [CrossRef]
Figure 1. The crystal structure of K2XF6 (X = Mn, Si, or Ge) with (a) Cubic structure and space group F m 3 ¯ m , (b) Rhombohedral structure with space group P 3 ¯ m 1 , and (c) Hexagonal structure with space group P 63 m c as seen from the c axis. The seven atoms represent clusters with (d) Oh, (e) D3d, and (f) C3v symmetry, with the Mn4+ ion in the core.
Figure 1. The crystal structure of K2XF6 (X = Mn, Si, or Ge) with (a) Cubic structure and space group F m 3 ¯ m , (b) Rhombohedral structure with space group P 3 ¯ m 1 , and (c) Hexagonal structure with space group P 63 m c as seen from the c axis. The seven atoms represent clusters with (d) Oh, (e) D3d, and (f) C3v symmetry, with the Mn4+ ion in the core.
Materials 16 04046 g001
Figure 2. Molecular orbital energies of Oh-K2MnF6, C3v-K2MnF6, Oh-K2SiF6, C3v-K2SiF6, D3d-K2GeF6, and C3v-K2GeF6 doped with Mn4+. The Valence Band (VB) is represented by black solid lines. The Conduction Band (CB) is shown by the black dashed lines. The t2g levels are indicated by the solid red lines, while the eg levels are indicated by the dashed blue lines.
Figure 2. Molecular orbital energies of Oh-K2MnF6, C3v-K2MnF6, Oh-K2SiF6, C3v-K2SiF6, D3d-K2GeF6, and C3v-K2GeF6 doped with Mn4+. The Valence Band (VB) is represented by black solid lines. The Conduction Band (CB) is shown by the black dashed lines. The t2g levels are indicated by the solid red lines, while the eg levels are indicated by the dashed blue lines.
Materials 16 04046 g002
Figure 3. Pure K2MnF6 and K2SiF6: Mn4+ multiplet energy diagrams. Additionally, demonstrated is the impact of CDC, CC, and lattice relaxation. The left side of each column explains the calculation using Oh-symmetric clusters, while the right side describes the calculation using C3v-symmetric clusters. Black and red lines denote the doublet and quartet states, respectively. When the lower symmetry (C3v) was used, these states were further divided into the a (dashed lines) and e (solid lines) categories. There are the doublet states 2E,.2T2, and 2T1, as well as the quartet states 4T2 and 4T1a. The 4A2 is the ground state. The absorption occurred during the electronic transitions from the ground 4A2 state to 4T2 and 4T1a states (U- and Y-band, respectively), as illustrated by the green and blue arrows. The emission, on the other hand, happened as an electronic transition from the 2E state to the ground 4A2 state (R-line), as illustrated by the red arrow. More information can be found in the text.
Pure K2MnF6 and K2SiF6: Mn4+ multiplet energy diagrams. Additionally, demonstrated is the impact of CDC, CC, and lattice relaxation. The left side of each column explains the calculation using Oh-symmetric clusters, while the right side describes the calculation using C3v-symmetric clusters. Black and red lines denote the doublet and quartet states, respectively. When the lower symmetry (C3v) was used, these states were further divided into the a (dashed lines) and e (solid lines) categories. There are the doublet states 2E,.2T2, and 2T1, as well as the quartet states 4T2 and 4T1a. The 4A2 is the ground state. The absorption occurred during the electronic transitions from the ground 4A2 state to 4T2 and 4T1a states (U- and Y-band, respectively), as illustrated by the green and blue arrows. The emission, on the other hand, happened as an electronic transition from the 2E state to the ground 4A2 state (R-line), as illustrated by the red arrow. More information can be found in the text.
Figure 3. Pure K2MnF6 and K2SiF6: Mn4+ multiplet energy diagrams. Additionally, demonstrated is the impact of CDC, CC, and lattice relaxation. The left side of each column explains the calculation using Oh-symmetric clusters, while the right side describes the calculation using C3v-symmetric clusters. Black and red lines denote the doublet and quartet states, respectively. When the lower symmetry (C3v) was used, these states were further divided into the a (dashed lines) and e (solid lines) categories. There are the doublet states 2E,.2T2, and 2T1, as well as the quartet states 4T2 and 4T1a. The 4A2 is the ground state. The absorption occurred during the electronic transitions from the ground 4A2 state to 4T2 and 4T1a states (U- and Y-band, respectively), as illustrated by the green and blue arrows. The emission, on the other hand, happened as an electronic transition from the 2E state to the ground 4A2 state (R-line), as illustrated by the red arrow. More information can be found in the text.
Pure K2MnF6 and K2SiF6: Mn4+ multiplet energy diagrams. Additionally, demonstrated is the impact of CDC, CC, and lattice relaxation. The left side of each column explains the calculation using Oh-symmetric clusters, while the right side describes the calculation using C3v-symmetric clusters. Black and red lines denote the doublet and quartet states, respectively. When the lower symmetry (C3v) was used, these states were further divided into the a (dashed lines) and e (solid lines) categories. There are the doublet states 2E,.2T2, and 2T1, as well as the quartet states 4T2 and 4T1a. The 4A2 is the ground state. The absorption occurred during the electronic transitions from the ground 4A2 state to 4T2 and 4T1a states (U- and Y-band, respectively), as illustrated by the green and blue arrows. The emission, on the other hand, happened as an electronic transition from the 2E state to the ground 4A2 state (R-line), as illustrated by the red arrow. More information can be found in the text.
Materials 16 04046 g003
Figure 4. K2GeF6: Mn4+ multiplet energy diagrams. Additionally demonstrated is the impact of corrections, including CDC, CC, and lattice relaxation. A calculation using clusters with D3d symmetry is described on the left side of each column, while a calculation using clusters with C3v symmetry is described on the right side. The Oh symmetry notations, in this instance, were borrowed. Black and red lines denote the doublet and quintet states, respectively; dashed (a level) and solid lines (e level) denote the multiplet splitting. There are the doublet states 2E,.2T2, and 2T1, as well as the quartet states 4T2 and 4T1a. The 4A2 is the ground state. The absorption occurred during the electronic transitions from the ground 4A2 state to 4T2 and 4T1a states (U- and Y-band, respectively), as illustrated by the green and blue arrows. The emission, on the other hand, happened as an electronic transition from the 2E state to the ground 4A2 state (R-line), as illustrated by the red arrow. More information can be found in the text.
Figure 4. K2GeF6: Mn4+ multiplet energy diagrams. Additionally demonstrated is the impact of corrections, including CDC, CC, and lattice relaxation. A calculation using clusters with D3d symmetry is described on the left side of each column, while a calculation using clusters with C3v symmetry is described on the right side. The Oh symmetry notations, in this instance, were borrowed. Black and red lines denote the doublet and quintet states, respectively; dashed (a level) and solid lines (e level) denote the multiplet splitting. There are the doublet states 2E,.2T2, and 2T1, as well as the quartet states 4T2 and 4T1a. The 4A2 is the ground state. The absorption occurred during the electronic transitions from the ground 4A2 state to 4T2 and 4T1a states (U- and Y-band, respectively), as illustrated by the green and blue arrows. The emission, on the other hand, happened as an electronic transition from the 2E state to the ground 4A2 state (R-line), as illustrated by the red arrow. More information can be found in the text.
Materials 16 04046 g004
Table 1. Mn-F Bond lengths (Å) of Oh-K2MnF6, C3v-K2MnF6, Oh-K2SiF6, C3v-K2SiF6, D3d-K2GeF6, and C3v-K2GeF6 doped with Mn4+.
Table 1. Mn-F Bond lengths (Å) of Oh-K2MnF6, C3v-K2MnF6, Oh-K2SiF6, C3v-K2SiF6, D3d-K2GeF6, and C3v-K2GeF6 doped with Mn4+.
K2XF6 Crystalsd1d2d3d4d5d6
Without relaxation
Oh-K2MnF62.0059202.0059202.0059202.0059202.0059202.005920
C3v-K2MnF61.7853111.7853111.7853141.7922691.7922711.792271
Oh-K2SiF6: Mn4+1.6829201.6829201.6829201.6829201.6829201.682920
C3v-K2SiF6: Mn4+1.6805381.6805421.6805421.6884631.6884631.688461
D3d-K2GeF6: Mn4+1.7702841.7702801.7702801.7702841.7702841.770284
C3v-K2GeF6: Mn4+1.7773601.7773571.7773571.8055701.8055701.805564
With relaxation
Oh-K2MnF62.0059202.0059202.0059202.0059202.0059202.005920
C3v-K2MnF61.7853111.7853111.7853141.7922691.7922711.792271
Oh-K2SiF6: Mn4+1.8070001.8070001.8070001.8070001.8070001.807000
C3v-K2SiF6: Mn4+1.8027441.8027531.8027531.8112481.8112481.811247
D3d-K2GeF6: Mn4+1.8100001.8099971.8099971.8100001.8100001.810000
C3v-K2GeF6: Mn4+1.7957491.7957501.7957501.8242501.8242501.824244
Table 2. Using the MnF62− model clusters with Oh, D3d, and C3v symmetry, Coulomb integral (eV) for the pure TM−3d atomic orbitals ( J AO ) and the molecular orbitals ( J MO ) were calculated. The adjustments were contrasted, including those with and without lattice relaxation. The orbital deformation parameter ( λ ) and the correlation correction factor (c) were multiplied by J AO to calculate the effective Coulomb integrals ( J eff ).
Table 2. Using the MnF62− model clusters with Oh, D3d, and C3v symmetry, Coulomb integral (eV) for the pure TM−3d atomic orbitals ( J AO ) and the molecular orbitals ( J MO ) were calculated. The adjustments were contrasted, including those with and without lattice relaxation. The orbital deformation parameter ( λ ) and the correlation correction factor (c) were multiplied by J AO to calculate the effective Coulomb integrals ( J eff ).
CompoundK2MnF6K2MnF6: Mn4+K2SiF6: Mn4+
Relaxed
K2GeF6: Mn4+ K2GeF6: Mn4+
Relaxed
SymmetryOhC3vOhC3vOhC3vD3dC3vD3dC3v
J AO 23.9224.3024.3624.1224.2123.8824.3524.2924.3124.26
J MO t 2 g 19.1919.9620.4019.3019.9118.3520.0719.9519.9119.88
J MO e g 16.5018.4319.0818.9118.0917.9318.6618.3218.3818.18
λ t 2 g 0.800.820.840.800.820.770.820.820.820.82
λ e g 0.690.760.780.780.750.750.770.750.760.75
c factor1.010.840.780.770.850.840.830.850.860.86
c λ t 2 g 0.810.690.650.610.700.640.690.690.700.70
c λ t 2 g 0.700.640.610.600.640.630.640.640.650.64
J eff t 2 g 19.4216.8115.8614.7716.9415.3916.6916.8717.0717.05
J eff e g 16.7015.5214.8314.4715.3915.0415.5115.4915.7515.60
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Novita, M.; Ristanto, S.; Saptaningrum, E.; Supriyadi, S.; Marlina, D.; Rondonuwu, F.S.; Chauhan, A.S.; Walker, B.; Ogasawara, K.; Piasecki, M.; et al. Study on Local-Structure Symmetrization of K2XF6 Crystals Doped with Mn4+ Ions by First-Principles Calculations. Materials 2023, 16, 4046. https://doi.org/10.3390/ma16114046

AMA Style

Novita M, Ristanto S, Saptaningrum E, Supriyadi S, Marlina D, Rondonuwu FS, Chauhan AS, Walker B, Ogasawara K, Piasecki M, et al. Study on Local-Structure Symmetrization of K2XF6 Crystals Doped with Mn4+ Ions by First-Principles Calculations. Materials. 2023; 16(11):4046. https://doi.org/10.3390/ma16114046

Chicago/Turabian Style

Novita, Mega, Sigit Ristanto, Ernawati Saptaningrum, Slamet Supriyadi, Dian Marlina, Ferdy Semuel Rondonuwu, Alok Singh Chauhan, Benjamin Walker, Kazuyoshi Ogasawara, Michal Piasecki, and et al. 2023. "Study on Local-Structure Symmetrization of K2XF6 Crystals Doped with Mn4+ Ions by First-Principles Calculations" Materials 16, no. 11: 4046. https://doi.org/10.3390/ma16114046

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