Next Article in Journal
Mechanism of Reduced Sintering Temperature of Al2O3–ZrO2 Nanocomposites Obtained by Microwave Hydrothermal Synthesis
Previous Article in Journal
Plasmonic Nanowires for Wide Wavelength Range Molecular Sensing
Previous Article in Special Issue
Nanostructure and Corresponding Quenching Efficiency of Fluorescent DNA Probes
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Characterization of Luminescent Materials with 151Eu Mössbauer Spectroscopy

1
Fraunhofer Application Center for Inorganic Phosphors, Branch Lab of Fraunhofer Institute for Microstructure of Materials and Systems IMWS, Lübecker Ring 2, 59494 Soest, Germany
2
University of Tennessee Space Institute, Tullahoma, TN 37388, USA
3
Faculty of Electrical Engineering, South Westphalia University of Applied Sciences, Lübecker Ring 2, 59494 Soest, Germany
*
Author to whom correspondence should be addressed.
Materials 2018, 11(5), 828; https://doi.org/10.3390/ma11050828
Submission received: 15 February 2018 / Revised: 2 May 2018 / Accepted: 9 May 2018 / Published: 17 May 2018

Abstract

:
The application of Mössbauer spectroscopy to luminescent materials is described. Many solids doped with europium are luminescent, i.e., when irradiated with light they emit light of a longer wavelength. These materials therefore have practical applications in tuning the light output of devices like light emitting diodes. The optical properties are very different for the two possible valence states Eu 2 + and Eu 3 + , the former producing ultraviolet/visible light that shifts from violet to red depending on the host and the latter red light, so it is important to have a knowledge of their behavior in a sample environment. Photoluminescence spectra cannot give a quantitative analysis of Eu 2 + and Eu 3 + ions. Mössbauer spectroscopy, however, is more powerful and gives a separate spectrum for each oxidation state enabling the relative amount present to be estimated. The oxidation state can be identified from its isomer shift which is between 12 and 15 mm/s for Eu 2 + compared to around 0 mm/s for Eu 3 + . Furthermore, within each oxidation state, there are changes depending on the ligands attached to the europium: the shift is more positive for increased covalency of the bonding ligand X, or Eu concentration, and decreases for increasing Eu–X bond length.

1. Introduction

In the last half of the 20th century, solid state physics has been responsible for a remarkable revolution in electronics with the replacement of the vacuum tube by semiconductor transistors. Not quite so dramatic but more visible, literally, is the replacement of incandescent lighting by light emitting diodes (LEDs) with luminescent materials. Among these materials are the lanthanides, which have unique magnetic, luminescent, and electrochemical properties and are therefore used for many different applications, such as magnets, batteries, superconductors or in optical devices. In particular, europium has attracted much attention due to its red emission in the trivalent state (Eu 3 + ), which is widely used in fluorescent lamps, displays, and recently for solid-state lighting [1,2]. The broad emission of divalent europium (Eu 2 + ) is used in lighting applications [3] as well as euro bank notes [4].
As a rule when substituting for trivalent ions (Al 3 + , Y 3 + , La 3 + ) europium generally forms Eu 3 + , while for divalent ions (Mg 2 + , Ca 2 + , Sr 2 + , Ba 2 + ), it goes in as Eu 2 + . In some cases, the optical activation with divalent europium turns out to be difficult, since Eu 2 + is easily oxidized to Eu 3 + . An exact knowledge of the Eu 2 + and Eu 3 + content is necessary to guarantee an efficient light output and to optimize the material for practical applications. Eu 2 + and Eu 3 + are clearly distinguishable by photoluminescence spectroscopy. Eu 2 + is characterized by a broad emission, while Eu 3 + shows narrow emission lines. Both Eu 2 + and Eu 3 + are sensitive to changes in the surrounding crystal field. Photoluminescence spectroscopy, however, does not allow a quantitative analysis of Eu 2 + and Eu 3 + ions. For example, Eu 2 + does not fluoresce in the fluorozirconate base glass, although it is clear from electron paramagnetic resonance (EPR) [5] and Mössbauer studies [6] that it is present in the glass.
X-ray absorption spectroscopy is one method to determine not only the charge of the doped europium ions, but also allows a quantitative analysis [7,8]. However, it requires access to synchrotron radiation sources.
151 Eu Mössbauer spectroscopy has been established as a sensitive tool to distinguish quantitatively between Eu 2 + and Eu 3 + ions and to investigate the local structure around europium ions in solids. Unlike electron paramagnetic resonance (EPR) which cannot be used for trivalent europium ions (since the 7 F 0 ground state is not paramagnetic), information about both oxidation states emerges directly from the Mössbauer isomer shift.
The Mössbauer Effect is the recoilless resonant fluorescence of gamma-radiation. It was discovered in 191 Ir by Rudolf L. Mößbauer in 1958 [9] for which he received the Nobel Prize in 1961. After it was observed in 57 Fe, the field developed so fast that the first International Mössbauer Conference took place in 1960 at the University of Illinois [10]. Since then, the Mössbauer Effect was found in many isotopes including 119 Sn, 121 Sb, and 151 Eu and has been applied in many fields of sciences, such as physics, chemistry, biology, and medicine. A review of 151 Eu work has been written by Grandjean and Long [11]. An index of Mössbauer data is available from the Mössbauer Effect Data Center [12].
Many review papers assessing Mössbauer spectroscopy as a characterization technique for a wide variety of materials have been published. A general review paper has been published in 1968 [13], followed by Mössbauer spectroscopy on iron [14], glass [15,16,17], Fe/S proteins [18], and europium chalcogenides [19] among others. No review paper is known for the investigation of europium luminescence using Mössbauer spectroscopy. In this article, its application to europium-containing inorganic luminescent materials is reviewed.

2. The Lanthanide Ions Eu 2 + /Eu 3 + and Their Optical Properties

Europium is the chemical element with the atomic number 63 and the electron configuration [Xe] 4 f 7 6 s 2 . As for all lanthanides, the most stable oxidation state is +3, but europium forms divalent compounds as well. Divalent europium is rather unstable and oxidizes in air to form trivalent Eu compounds. Both Eu 2 + and Eu 3 + have an incompletely filled 4 f shell, which is shielded by the filled 5 s and 5 p shells. This peculiar electronic configuration is responsible for their unique optical and luminescent behaviour. The energy level diagrams for Eu 2 + and Eu 3 + are shown in Figure 1. Due to the shielding, 4 f n energy levels are only weakly influenced by the host material and can be depicted as solid lines each with a characteristic energy. Using the term symbol 2 S + 1 L J where S is the total spin quantum number, L is the total orbital quantum number, and J is the total angular momentum quantum number the ground state configuration for Eu 2 + and Eu 3 + are 8 S 7 / 2 and 7 F 0 , respectively.
Intra-configurational 4 f 4 f transitions are forbidden by the parity selection rule. However, the 4 f wave function mixes with an opposite parity wave function, e.g., 5 d , and the 4 f n transitions gain intensity. The inter-configurational 4 f 5 d transitions are allowed and appear in the spectra as broad absorption and emission bands; they are depicted as grey bands in Figure 1. Additionally, their energy depends strongly on the composition of the host structure.
Whereas the red luminescence of trivalent europium results from forbidden 4 f 4 f transitions, the luminescence of divalent europium results from allowed 4 f 5 d transitions. Both are discussed in more detail in the following sections.

2.1. Optical Properties of Eu 2 +

2.1.1. Eu 2 + in Luminescent Fluoride and Oxide Glasses

ZBLAN glasses were discovered accidentally in 1975 by Poulain and Lucas [23] at the University of Rennes in France. They were typically used for optical fibers due to their high infrared transmittance [24,25]. ZBLAN glass, however, is fragile and sensitive to acids. Interestingly, its manufacture has been initiated on the International Space Station to avoid defects [26]. In 2003, Schweizer et al. thought of substituting some of the BaF 2 for BaCl 2 and adding Eu 2 + [27]. On heating, the BaCl 2 drops out as a crystallite, incorporating the optically active divalent europium. This initiated an investigation into their suitability as X-ray image plates for medical diagnosis [28]. Further applications of fluorozirconate glass are optical devices, such as colour displays [29] or up-converting and down-converting glass layers for solar cells [30,31]. Glass ceramics with hexagonal phase crystallites can be used as scintillators [32] while the orthorhombic phase is suitable for storage phosphors [33].
Europium can be incorporated in its divalent and trivalent state in ZBLAN glass. Eu 3 + shows its typical emissions in the red spectral range (see Section 2.2) in the fluorozirconate base glass. Upon annealing, some of the Eu 2 + ions are incorporated into the BaCl 2 nanocrystals leading to an intense Eu 2 + -related fluorescence under ultraviolet excitation. After 20 min of annealing at 260 C the Eu 2 + fluorescence spectrum shows a main emission band at 407 nm (Figure 2, solid curve) and a weaker, but broader emission at 485 nm. After annealing at 290 C (Figure 2, dashed curve), the 407 nm emission is shifted to 402 nm, while the 485 nm band has completely disappeared. The 402-nm and 407-nm emission bands are attributed to Eu 2 + in hexagonal and orthorhombic BaCl 2 , respectively, while the origin of the additional weaker but broader band at 485 nm is unknown. These observations are described in detail in [34]. The quantum efficiency spectrum of Eu 2 + in ZBLAN containing orthorhombic BaCl 2 crystallites is shown on the left.
Oxide glasses containing divalent lanthanide ions have attracted considerable attention as optical and magneto-optical devices. These glasses are advantageous for applications, such as frequency and time domain optical memories. Aluminoborate glasses containing a large amount of Eu 2 + show a Faraday effect [35]. The paramagnetic Faraday effect can be used for magneto-optical devices, such as optical isolators, optical switches, and optical shutters [36].

2.1.2. Eu 2 + in (Persistent) Phosphors/Aluminates

Persistent luminescence (also known as phosphorescence) is a phenomenon in which light (UV, visible or IR) is emitted for minutes, hours or even days after the initial excitation. The mechanism underlying this phenomenon is not fully understood but is known to involve energy traps. These traps are filled during excitation. After excitation, the stored energy is released to emitter centres, which gradually emit the light. There are many persistent phosphors, but the most studied are the strontium aluminates doped with Eu 2 + and Dy 3 + due to their high brightness, long lifetime, and stability. The slow decay gives the ion the ability to store information, which may be read later by optical (laser) or thermal stimulation [37]. However, large-scale production of these materials needs to be developed for them to realize their full commercial potential. For a detailed review of persistent phosphors, the reader should look at the following references [38,39].
BaMgAl 10 O 17 :Eu 2 + has two broad excitation bands at approximately 270 nm and 300 nm and shows a blue luminescence with the peak near 450 nm. Together with a red-emitting phosphor and a green-emitting phosphor, it yields a white emitting blend for fluorescent lamps and plasma display panels [40,41]. However, it is unstable in a variety of lamp-related processing conditions and also during the lamp life.
Eu 2 + in other aluminates and alumosilicates shows persistent luminescence, e.g., CaAl 2 O 4 :Eu 2 + (blue) [42], SrAl 2 O 4 :Eu 2 + (green) [43], Sr 4 Al 14 O 25 :Eu 2 + (blue) [44], and CaAl 2 Si 2 O 8 :Eu 2 + (white) [45]. These compounds have applications to thermoluminescence, used for dating and radiation monitoring [46].

2.1.3. Eu 2 + in Other Luminescent Materials

CaS-based phosphors are studied for dosimetry applications [47]. For Eu 2 + -doped CaS containing additional impurities (such as Sm 2 + ), the stored dose can be read out upon optical stimulation in the infrared spectral range (in a so-called optically stimulated luminescence (OSL) process) instead of thermoluminescence [48].
Nitridosilicates provide efficient luminescent materials that are industrially applied in commercial phosphor-converted LEDs [49,50]. Here, Eu 2 SiN 3 is of particular interest since it is the only mixed-valence europium nitridosilicate, i.e., it has two different crystallographic Eu sites at which one site is occupied by Eu 2 + ions, while the other site is occupied by Eu 3 + ions [51]. It is also noteworthy that Eu 2 SiN 3 has a black colour due to its small band gap of 0.2 eV [50].
Oxonitridosilicates combine structural features and properties of both, oxosilicates and nitridosilicates. EuSi 2 O 2 N 2 shows a narrow yellow emission and is investigated for 2- and 3-phosphor-converted light-emitting diodes [52].

2.2. Optical Properties of Eu 3 +

Eu 3 + shows in different materials an efficient red luminescence with high quantum efficiencies. In 1989, the quantum efficiency of Eu 3 + in BaCa 2 Y 6 O 12 was found to be up to 25% [53]. The quantum efficiency of nanocrystalline powder of Lu 2 O 3 :Eu 3 + reaches 90% [54] and that of bulk Y 2 O 3 :Eu 3 + of 92% [55]. In borate and in fluorozirconate glass, the Eu 3 + quantum efficiency was determined to be 86% [56] and 94% [57], respectively.

2.2.1. Eu 3 + in Luminescent Glasses

Eu 3 + was used in fibre lasers in the early 1990s, mainly as a co-dopant with other rare earths to enhance efficiency [58,59]. More recently, it has been used as a down-converter in photovoltaic application [60,61] and to produce white light in LED’s, also as a co-dopant [56,57,62]. Energy transfer is the key to increase efficiency in both cases, the details of which are discussed within the references given here. The literature on Eu 3 + -doping of borate glasses is more plentiful than that involving ZBLAN. Again, recent research has focused on the creation of white light and many glasses are not pure borates but combinations, such as fluoroborate, borogermanate, aluminoborate, lead borate, and borosilicate, for example. Whatever the host, Eu 3 + is responsible for the red component of emission and other rare earths complement Eu 3 + in order to emit other colours, although white light is the most desirable right now. Older papers, starting in the early 1990s, focused on the study of fundamental luminescence properties.
The absolute PL quantum efficiency (QE) and PL emission spectra of Eu 3 + -doped borate (66B 2 O 3 ·33BaO·1Eu 2 O 3 , values in mol %) and ZBLAN glass (51ZrF 4 ·20BaF 2 ·20NaF·3.5LaF 3 ·3AlF 3 ·0.5InF 3 ·2EuF 3 , values in mol %) are depicted in Figure 3. The maximum quantum efficiency value is found in ZBLAN glass and amounts to 94% for 395-nm excitation ( 7 F 0 to 5 L 6 transition). The glasses can also be excited in the blue spectral range, for instance at 465 nm ( 7 F 0 to 5 D 2 transition), resulting in QE values of 70% and 82% for ZBLAN and borate glass, respectively. In comparison to ZBLAN glass, borate glass provides a higher QE in the longer wavelength range, but a lower one in the short wavelength range.
In both glass systems, transitions from the excited state, 5 D 0 , to the ground state levels, 7 F J ( J = 1 , 2 , 3 , 4 , 5 , and 6) are observed leading to the typical emission in the red spectral range. In ZBLAN glass, additional emissions in the ultraviolet/blue spectral range are observed originating from the excited states, 5 D 1 and 5 D 2 , to the ground states, 7 F J ( J = 0 , 1 , 2 , and 4). In borate glass, emissions from the excited states, 5 D 1 and 5 D 2 , are quenched due non-radiative relaxation, whereas in ZBLAN glass the probability for radiative emissions from these levels is significantly higher. The 5 D 0 and 5 D 1 states are separated by 1750 cm 1 [20]. In the case of borate glass with a maximum phonon frequency of 1400 cm 1 [63,64], only one to two phonons are needed to bridge the gap, while, for ZBLAN glass, with a maximum phonon frequency of 580 cm 1 [65], more than three phonons are necessary. Thus, the non-radiative transition rates in ZBLAN glass are significantly smaller than in borate glass enabling radiative emission.
The electric-dipole transition 5 D 0 to 7 F 2 is hypersensitive to variations in crystal symmetry [1]. The high intensity of this transition in borate glass indicates the amorphous nature of the matrix material without inversion symmetry for the Eu 3 + ion. For ZBLAN glass, the intensity of the 5 D 0 to 7 F 1 transition is higher than in borate glass, which implies a higher crystallinity of ZBLAN glass compared to borate glass. Therefore, Eu 3 + is a useful spectroscopic probe of the environment surrounding the lanthanide ion.

2.2.2. Eu 3 + in Lanthanide Oxides

Eu-doped sesquioxides Ln 2 O 3 (Ln = In, Sc, Y, La, Gd, Lu) are very important materials as they are used as the red-emitting phosphors in fluorescent lamps and colour television projection tubes [66]. Y 2 O 3 :Eu 3 + nanoparticles in spherical morphology are used for flat-panel displays [67]. Lu 2 O 3 :Eu is a very attractive host for scintillators [68] or x-ray phosphors [69], due to a high density of lutetia, which ensures that ionizing radiation is efficiently absorbed in relatively thin layers of lutetia-based phosphors.

2.2.3. Eu 3 + in Other Luminescent Materials

The production of III-V semiconductor based LEDs with efficient emission in the green and red spectral range is still challenging. GaN, doped with erbium and europium, enables the emission in the green and red spectral range, respectively, for lighting applications of optoelectronic devices [70].
Pyrochlore materials are known for their good thermal properties. Lanthanide-doped pyrochlores are used for contact-free surface temperature measurements [71].
Europium-doped titanium dioxide is developed as a substitute for the high-cost red-emitting phosphor Y 2 O 3 :Eu. TiO 2 :Eu is demonstrated to be a good sensitizer to absorb light and transfer energy to Eu 3 + ions [72,73]. It is also advantageous for practical applications due to its low cost, its chemical and thermal stability, and its good mechanical properties [72,74].
YVO 4 :Eu is a well-known red phosphor applied in cathode ray tubes, fluorescent lamps, and plasma displays. It provides high efficiency, colour purity, and thermal stability [75,76]. In addition, it is used as a fluorescent biological label for the detection of Na + channel dynamics on cell membranes [77]. EuVO 4 might be suitable to track biological systems, such as histidine and bovine serum albumin [78].
Zirconia is a widely-used material in optics due to its wide band-gap, high transparency, high refractive index and hardness [79]. Eu-doped ZrO 2 has been investigated for lamp and display applications [80].
Recently, Eu-doped yttrium aluminium garnet (Y 3 Al 5 O 12 , YAG) gained increasing attention as phosphor in white LEDs and laser-active material [81,82]. It crystallizes in the cubic form [81].

3. Europium Mössbauer Spectroscopy

3.1. Mössbauer Isotope 151Eu

Eu has two naturally occurring isotopes, namely 151 Eu (47.82%) and 153 Eu (52.18%) of which the former is the most useful for Mössbauer spectroscopy owing to its low γ -ray energy of 21.54 keV. The source used is 151 Sm, and its decay scheme is shown in Figure 4. The half-life of 151 Sm with 90 years is very long [83]. Two different β -decays occur. Only 0.9% decays into the excited state of 151 Eu. The remaining 99.1% goes to the ground state [83]. The 21.54 keV transition from the excited state with spin + 7 / 2 to the ground state of 151 Eu with spin + 5 / 2 has a lifetime of 14 ns and a resulting linewidth of 47 neV. Mössbauer isotopes must have a long lifetime for the decay of the excited state and very low lying excited states. These criteria exclude some isotopes. 151 Eu is the most used Mössbauer isotope of the lanthanide elements.
To produce a spectrum, the γ -ray energy is varied using the Doppler effect by moving the 151 Sm source relative to the 151 Eu-containing absorber. A plot of transmitted γ -ray counts against velocity yields the Mössbauer spectrum. The velocities needed are of the order of millimetres per second (in non-SI units this is a furlong per fortnight—a snail’s pace!). The linewidth is governed by the lifetime of the 21.54 keV transition and is 2.52 mm/s. The 21.54 keV radiation may also be produced with synchrotron radiation with an appropriate monochromator. Such facilities can be found at the European Synchrotron Radiation Facility (ESRF) in Grenoble (France), the Applied Photon Source (APS) at the Argonne National Laboratory (USA) or at the Super Photon ring (SPring-8) at the Harima Science Park (Japan).
Mössbauer measurements are usually performed in transmission geometry, but using backscattering geometry, γ -rays, x-rays, conversion electrons or Auger electrons may be detected as the excited nuclear state decays back to the ground state. For the 151 Eu isotope, there are no conversion electrons [84,85], but Bibicu et al. [85] found that this isotope emits Auger electrons. The transmission geometry gives volume information, while the backscattering geometry gives predominantly surface information.
In Mössbauer spectra, three types of nuclear interactions are observable: (i) isomer (chemical) shift; (ii) quadrupole splitting; and (iii) magnetic (hyperfine) splitting.

3.2. Isomer Shift

The isomer shift arises from the electric monopole interaction between the nucleus and the surrounding electron shell. The overlap of the charge density distribution of s-electrons with the nucleus causes a shift of the energy levels of the nucleus. The shift is observable because the ground and excited nuclear states have different radii and hence different overlaps with the electron cloud. It cannot be measured directly and therefore is quoted relative to a known absorber. In this paper, the values are given with respect to EuF 3 . The isomer shift is a good indicator for the valence state of the isotope. It is useful for the investigation of valence states, ligand bonding states, and electron shielding.
Eu 3 + compounds exhibit isomer shifts between 0 and + 3 mm/s. The large difference in isomer shifts of about 12 mm/s between divalent and trivalent Eu compounds results mainly from the shielding effect of the additional 4 f electron in Eu 2 + compounds; Eu 2 + has the configuration 4 f 7 , while Eu 3 + has 4 f 6 . Other effects like bond lengths, covalency, and coordination numbers produce less pronounced variations up to ± 2 mm/s for Eu 3 + and ± 1 mm/s for Eu 2 + but are clearly resolved due to the large difference between the nuclear radii of the ground and the excited state for the 151 Eu resonance, respectively. The isomer shifts of Eu in binary compounds are listed in Table 1. It is evident from Table 1 that fluorine is the most ionic ligand and therefore it is mentioned first in the following chapters.

3.3. Quadrupole Interaction

Quadrupole splitting may arise for nuclei in states with an angular momentum quantum number I > 1 / 2 due to their non-spherical charge distribution so that they have an electric quadrupole moment Q. This causes a split of the nuclear energy levels in the presence of an anisotropic electric field (anisotropic electronic charge distribution or ligand arrangement) when the lattice has a non-cubic structure. The charge distribution is characterized by the electric field gradient (EFG). From the quadrupole splitting, information on oxidation state, spin state, and site symmetry can be obtained. In a compound with cubic symmetry, the electric field gradient (and therefore the quadrupole interaction parameter) is zero and a single transition is observed [11], which is the case for the majority of the papers, summarized in this review article. If a threefold or fourfold axis is present, the electronic charge distribution will be symmetrical and eight transitions are allowed (see Figure 5). If there is no threefold or fourfold symmetry axis passing through the nucleus, the components of the electric field along the principal axes are different, the asymmetry parameter is non-zero, and there are 12 allowed transitions [86].
The quadrupole splitting is usually only partly resolved owing to the linewidth of about 2.5 mm/s. This gives an asymmetrically broadened line profile (see Figure 5), which for small interactions, has to be distinguished from overlapping contributions from more than one inequivalent site.

3.4. Magnetic Hyperfine Interaction

Magnetic hyperfine splitting is caused by the interaction between the nuclear magnetic moments and the magnetic field of the electrons on the atom. It is only observed for Eu 2 + and to observe it requires magnetic dilution so that electron spin relaxation rates are slow. The paramagnetic hyperfine fields are typically of the order of 30 T.
The nuclear spin I splits into 2 I + 1 sublevels, i.e., six levels for the ground state of 151 Eu ( I = 5 / 2 ) and eight for the excited state ( I = 7 / 2 ). The selection rules Δ m = 0 , ± 1 give rise to a symmetric 18-line spectrum, where, because of the smaller g-value of the excited state, it has the appearance of a six-line spectrum with broadened lines (see Figure 6).

3.5. Linewidth (FWHM)

The linewidth of the absorption line is dependent on the lifetime of the nuclear excited state [89]. In the case of an ideally thin absorber, the linewidth is twice the natural linewidth [89]. Usually, the linewidth of an amorphous material is broader than that of the corresponding crystalline material, e.g., Eu 3 + in metaphosphate has a linewidth of 1.96 mm/s and 1.76 mm/s for the amorphous and crystalline material, respectively [90].

4. Materials Overview

An overview of different luminescent materials, containing Eu 2 + and Eu 3 + with the corresponding isomer shifts at room temperature, is given in Table 2 and Table 3, respectively. The isomer shift is in the range 9.7 to 14.3 mm/s for Eu 2 + and 0.93 to + 1.47 mm/s for Eu 3 + .
The isomer shifts listed in Table 2 and Table 3 are visualized in Figure 7, sorted for different material systems. Fluorides show isomer shifts from 14.3 to 13 mm/s and oxide glasses from 13.5 to 13 mm/s. Eu 2 + in different kinds of aluminates have the largest range of isomer shifts from 9.7 to 15 mm/s. For materials, such as vanadates, sulphides, nitrides, and titanium dioxide, only a few papers were found and therefore depicted as “other” with isomer shifts ranging from 10.5 to 12.5 mm/s. For Eu 3 + , the isomer shifts vary from 1 to + 1.4 mm/s with fluorides having isomer shifts at approximately 0 mm/s, because the isomer shifts in this paper are given relative to EuF 3 . Eu 3 + has isomer shifts from 0 to 1 mm/s in oxide glasses and 0.6 to + 0.6 mm/s in vanadates. In aluminates and other materials, the Eu 3 + isomer shifts range from approximately 0.9 to + 1 mm/s.

4.1. Fluorides

CaF 2 with Eu 2 + impurities occurs in nature as the mineral fluorite (or fluorspar) and the term fluorescence originates from its luminescent properties. Eu 2 + in CaF 2 shows strong luminescence, while that of Eu 3 + is much weaker [128]. Implanted Eu 2 + substituting for Ca 2 + luminesces in the violet spectral range at 420 nm, while a second emission band at 680 nm arises from interstitial sites; the latter is eliminated on heating [129].
The 151 Eu Mössbauer spectra of highly diluted (0.1 mol %) Eu 2 + ions in CaF 2 showed an almost temperature-independent asymmetrically split pattern, arising from the paramagnetic hyperfine interaction AS.I in a cubic crystal field with slow electron spin relaxation. In a small external magnetic field B of 0.2 T such that g μ B B > A , an almost symmetrical pattern was observed. Both, the spectra with and without an external field, are well described using the spin Hamiltonian and previous electron paramagnetic resonance data. A more concentrated (2 mol % Eu 2 + ) sample exhibited a strongly broadened symmetrical resonance line due to an increased Eu–Eu spin relaxation rate in an external magnetic field of 0.2 T. The Mössbauer spectra exhibited further broadening and additional magnetic structures due to the reduced relaxation rate. When a large field of 6 T was applied such that g μ B B is much larger than the crystal field splitting, a fully resolved hyperfine pattern was observed at 2.5 K, with an effective field at the Eu nuclei of 33.7 T; at higher temperatures, superimposed patterns originating from excited electronic states were observed in the spectra [92].
Fluorophosphate glasses with the composition (45 x )AlF 3 ·xAlPO 4 ·5EuF 3 ·30CaF 2 ·20BaF 2 ( x = 0 –20) showed only trivalent Eu [107]. The isomer shift increases slightly with increasing Eu concentration.

4.2. Fluorochlorozirconate (FCZ) Glasses

Much work has been done on fluorochlorozirconate (FCZ) glasses and glass ceramics. Coey et al. [93] made measurements on 61ZrF 6 ·12BaF 2 ·7ThF 4 (values in mol %) doped with 20EuF 2 . Most of the europium was Eu 2 + with an isomer shift of 14.18 mm/s corresponding to a large coordination number ranging between 8 and 12 which is typical for glasses. While the Eu 2 + resonance line in the glass is extremely broad compared with EuF 2 , the Eu 3 + line is at least as narrow as that of EuF 3 . Measurements of the variation of the absorption at different temperature enable the relative binding strengths of Eu 2 + and Eu 3 + to be determined. The binding for Eu 2 + ( Θ Debye = 145 K) was weaker than that of Eu 3 + ( Θ Debye = 261 K).
An important material in computed radiography is ZBLAN, which contains 51ZrF 4 ·17BaF 2 ·3.5LaF 3 ·3AlF 3 ·20NaF (values in mol %), which, when doped with Eu 2 + , has applications as storage phosphors [33]. Weber et al. [91] showed that EuCl 3 heated at 710 C for 10 min resulted in almost equal amounts of EuCl 2 and EuCl 3 . This could be important for production of EuCl 2 storage phosphors since EuCl 2 is more expensive than EuCl 3 . They also have shown that samples made with 5 mol % EuCl 2 contained both EuCl 2 and EuCl 3 in the ratio 78:22 (Figure 8a). Samples made with a mixture of 2.5 mol % EuCl 2 and 2.5 mol % EuCl 3 had a ratio 37:63 (Figure 8b), i.e., 13% of the Eu 2 + was oxidized. Johnson et al. [28] showed that EuCl 2 as raw material oxidizes to Eu 2 O 3 while EuCl 2 in the glass oxidizes to EuCl 3 . Pfau et al. [94] studied ZBLAN doped with 0.5 mol % InF 3 (to keep the ZrF 4 from being reduced) and with BaCl 2 and 5 mol % EuCl 2 substituted for BaF 2 . About 88% of the europium was in the divalent state. When 5 mol % EuF 2 was used, only 70% of the europium was in the divalent state. The isomer shift is not influenced by the thermal treatment of the glasses and amounts to approximately 14 mm/s. The Debye temperatures are Θ Debye = 147 K for Eu 2 + and Θ Debye = 186 K for Eu 3 + .

4.3. Aluminates

BaMgAl 10 O 17 (BAM) has the β -alumina structure and, when doped with Eu 2 + , it is a blue-emitting phosphor used for white light generation. It is generally unstable which is a problem for lighting applications. Fraknoy-Körös et al. [100] prepared Ba 0.9 Eu 0.1 MgAl 10 O 17 in air and Mössbauer measurements show only trivalent Eu. In a reducing atmosphere both Eu 3 + and Eu 2 + are obtained. Without Mg and in a reducing atmosphere, mainly Eu 2 + is produced. Boolchand et al. [98] investigated BAM containing different Eu concentrations of 6%, 12%, and 20%. The Mössbauer spectra were measured at 4.2 K and show significant magnetic hyperfine structure, though Boolchand interprets the spectra in terms of large quadrupole splitting. The isomer shifts are similar for all investigated Eu concentrations. Similar results are shown by Mishra et al. [130].
MAl 12 O 19 (M = Ca, Sr, Ba) has the magnetoplumbite structure (related to β -alumina). The Mössbauer measurements of Hintzen et al. [101] also revealed a small amount of Eu 3 + (<10%). The linewidth decreased with increasing Eu concentration from 11 mm/s to 3.6 mm/s, which is presumably due to unresolved paramagnetic hyperfine structure resulting from increased spin-spin relaxation. At a concentration of 25%, the emission spectrum shows an additional band at 490 nm as a consequence of the presence of very small traces of a second phase. No influence of the temperature on the spectra was observed in the range from 10 K to 300 K.
Arakawa et al. [95] reported an isomer shift of 9.7 mm/s which decreases with increasing size of the alkaline earth atom, a result of the decreased electron density at the Eu nuclei. The isomer shift decreases for increasing Eu concentration. The emission spectra of BaAl 12 O 19 and CaAl 12 O 19 contain one broad emission. For CaAl 12 O 19 , the peak shifts from 425 nm to 439 nm with an increase of Eu from 5% to 50% relative to Ca, while the peak of BaAl 12 O 19 remains at approximately 450 nm. SrAl 12 O 19 :Eu 2 + shows two emission peaks at 390 nm and 530 nm, for all investigated Eu concentrations.
Amorphous Al 2 O 3 doped with Eu 3 + shows a broad line, which was decomposed (not necessarily uniquely) into contributions from eight different quadrupole-split sites, as would be expected from an amorphous material [108]. The average isomer shift of 1.107 mm/s suggests strong covalent bonds and leads to the conclusion that Eu 3 + ions replace Al 3 + ions in the Al 2 O 3 matrix.
Hölsä et al. [42] investigated CaAl 2 O 4 with a low Eu concentration due to the segregation of the Eu ion from the CaAl 2 O 4 phase. With Eu 2 + doping, the compounds exhibit phosphorescence (persistent luminescence). Their Mössbauer spectra show a very broad line resulting from Eu 2 + but also a small amount of Eu 3 + . EuAl 2 O 4 has a green emission with the maximum at 515 nm [97]. Mössbauer spectra show two crystallographic sites for Eu 2 + with different isomer shifts of 13.4 mm/s and 12.83 mm/s. The photoluminescence excitation spectrum shows two broad bands at 380 nm and 430 nm.
Tronc et al. [99] made measurements on LaMgAl 11 O 19 , which has the magnetoplumbite structure for compositions with 100% and 30% Eu replacing La using different methods of preparation. The spectra showed line broadening arising from paramagnetic hyperfine splitting, which was partially resolved in the most dilute (30% Eu) sample.

4.4. Oxide Glasses

Tanabe et al. [114] found that in silicate and aluminate glasses the isomer shift of Eu 3 + increases with increasing optical basicity of the glasses. The optical basicity measures the electron donation by the oxygen anions to the metal ion used as a probe (Eu), the correlation indicates that the charge transferred to the Eu 3 + ion occupies mainly 6 s shells and hence the phosphates have the smallest charge in the 6 s shells compared to the other oxide glasses.
In silicate glasses M 3 Na 8 Si 13 O 65 (M = Mg, Ca, Ba) the Eu 3 + isomer shift decreases with decreasing modifier cation size, i.e., increasing electronegativity (Figure 9) [112]. The shift is close to that of Eu 2 O 3 [110], suggesting that Eu 3 + ions locate in sites similar to those in Eu 2 O 3 . Musić et al. [111] obtained similar values for the isomer shift of Eu 3 + in sodium borosilicate glasses.
In related aluminosilicate crystal Ba 0.95 Eu 0.05 Al 2 Si 2 O 8 , the europium is mostly Eu 2 + with an isomer shift of 14 mm/s. About 5 to 10% of the europium is Eu 3 + with an isomer shift close to 0 mm/s [45].
In metaphosphate glasses M 5 Al 4 P 25 O 76 (M = Mg, Ca, Ba), the Eu 3 + isomer shift increases with decreasing modifier cation size, i.e., increasing electronegativity. This is in contrast to the behaviour in silicate glasses [112] and is ascribed to the effect of the short π -bonds between the modifier atoms and the non-bridging phosphate chains. Metaphosphate glasses containing Zn, Sr, and Pb were investigated by Concas et al. [115]. The glass containing Pb has a lower isomer shift compared to the other two samples. A sodium phosphate glass containing Eu 3 + and Ce 3 + was irradiated by multipulse excimer-UV-laser irradiation, leading to an extinction of the luminescence [113]. The as-made glass showed an isomer shift of approximately 0.79 mm/s and decreases to 0.37 mm/s after irradiation. It is believed that the electronic traps created by irradiation expand the wave function of 5 s electrons, thus decreasing the electronic density at the nucleus.
The borate glasses B 2 O 3 :Eu and B 2 O 3 -A 2 O 3 :Eu (A = Li, Na, K) were studied by Winterer et al. [102] for Eu concentrations between 0.1% and 33%. The isomer shifts of 13.0 mm/s corresponded to Eu 2 + . The more dilute glasses showed paramagnetic hyperfine splitting with fields of 35 T. The more concentrated ones showed quadrupole broadened Eu 2 + lines, and also some Eu 3 + . Fujita et al. [36] obtained similar results for B 2 O 3 -Na 2 O:Eu glasses. In sodium borate glass prepared under a reducing atmosphere, Fujita [131] observed the Eu 2 + and Eu 3 + absorption bands in the Mössbauer spectra and investigated spectral hole burning in Eu 3 + emission spectra. The relative hole area increases with the absorption area ratio, R, of Eu 2 + /Eu 3 + in the Mössbauer spectrum up to a value of R = 0.3 and decreases slightly for larger ratios.
Fujita et al. [36] studied 15EuO·85(( 1 x )B 2 O 3 ·xNa 2 O) and found that the Eu 2 + isomer shifts increase as the concentration of Na 2 O increases and decrease as the concentration of EuO increases. In borate glasses with increasing network modifier content, the amount of three-coordinated boron decreases and that of four-coordinated boron increases up to about 33 mol % Na 2 O content, beyond which the four-coordinated boron starts to decrease (boron anomaly) due to the appearance of non-bridging oxygen. Hence, the electron density at the Eu nucleus drastically increases in this compositional region, since the non-bridging oxygen has more electron donation ability than the bridging oxygen and increases the 6 s electron density effectively.
Nemov et al. [126] investigated germanium glasses (BaGeO 3 ) 1 x y (Al 2 O 3 ) x (0.45CaF 2 ·0.55MgF 2 ) y with Eu 2 O 3 concentrations varying from 5 mol % to 20 mol %. Europium was observed only in the trivalent state. The influence of the Eu content was minimal, while the increase of the fluorine content led to a larger isomer shift as well as to a higher frequency (shorter wavelength) of the PL emission band. The linewidth of both the isomer shift and the emission band decreased for increasing fluorine content.

4.5. Lanthanides and Related Compounds Sc2O3, Y2O3, In2O3

The crystal structures of the lanthanide sesquioxides Ln 2 O 3 are complicated: cubic, monoclinic, and hexagonal structures have been observed. For cubic Eu 2 O 3 , shift values are close to 1.0 mm/s [36,117,132,133,134]. For Eu 3 + in lanthanide oxides and the related compounds Sc 2 O 3 , Y 2 O 3 , and In 2 O 3 the isomer shifts increase in the sequence Gd 2 O 3 , Y 2 O 3 , Sc 2 O 3 , Lu 2 O 3 , and In 2 O 3 , while the mean Ln-O distance decreases in the same sequence (Table 3) [66,119]. The monoclinic system also has two sites for the Ln ions, while the hexagonal system has only one site. Hintzen et al. [66] compared monoclinic and cubic Eu 2 O 3 and obtained slightly larger isomer shifts for the monoclinic structure.
The spectra of Sc 2 O 3 :Eu 3 + show a clear decomposition into two lines (Figure 10) [66,117]. Concas et al. [117] investigated cubic Sc 1.8 Eu 0.2 O 3 and found isomer shifts of 1.18 and 1.38 mm/s for the sites C 3 i and C 2 , respectively. Hintzen et al. [66] obtained a higher difference between the isomer shifts of the two different sites for cubic Sc 1.9 Eu 0.1 O 3 of 0.7 and 2.59 mm/s.
Bohus et al. [67] and Kuzmann et al. [116] investigated monodisperse Y 2 O 3 :Eu 3 + spherical nanoparticles as well as Y 2 O 3 @Eu 3 + core-shell structures. The core-shell nanoparticles had an isomer shift of 1.03 mm/s, which agrees with that found for pure Eu 2 O 3 and is independent of the Eu 3 + doping level. Other papers gave similar results for Y 2 O 3 :Eu 3 + [66,100,117].
Y 2 x Eu x WO 6 ( x = 0.05–0.4) was investigated by van Noort and Pompa [123]. The Eu ions were in the trivalent state and occupied three different sites, two sites with symmetry C 2 and one site with symmetry C 1 . Small differences in covalency produce the isomer shifts obtained of 1.25 mm/s, 0.65 mm/s, and 0.05 mm/s, respectively, and which increased with Eu 3 + concentration, due to an increase of the lattice parameter a (Figure 11).
Monoclinic and cubic Gd 1.9 Eu 0.1 O 3 showed different average isomer shifts of 1.08 and 1.05 mm/s, respectively, resulting from the different oxygen surroundings of the Eu 3 + ions in the two crystallographic structures [66,86].

4.6. Vanadates

Bibicu et al. [120] observed two broad lines in the Mössbauer spectrum of YVO 4 :Eu 3 + nanocrystals doped with 5 at.% showing the presence of Eu 3 + ions in different possible micro-environments. Fraknoy-Körös et al. [100] found similar results. Li et al. [121] investigated EuVO 4 prepared at two different temperatures resulting in different phases, which were identified as scheelite and zircon. The scheelite phase showed a strong luminescence. Stadtnik et al. [105] investigated three different vanadate compounds: Eu 2 VO 4 , Eu 3 V 2 O 7 , and Eu 0.6 Sr 0.4 VO 2.96 . The first two compounds showed Eu 2 + and Eu 3 + , while the latter showed only Eu 3 + . The amount of Eu 2 + was found to be 39% and 17% and had quadrupole interactions of 18.0 mm/s and 13.2 mm/s for Eu 2 VO 4 and Eu 3 V 2 O 7 , respectively. Divalent and trivalent Eu ions are found to exist simultaneously at the same crystallographic sites in Eu 2 VO 4 and Eu 3 V 2 O 7 .

4.7. Titanates

TiO 2 :Eu amorphous powder showed only Eu 3 + sites with isomer shifts of approximately 0.5 mm/s independent of Eu concentration, while the linewidth increases with increasing Eu concentration [125]. Calcination at temperatures above 500 C lead to the formation of brookite and anatase phases, while for temperatures higher than 1000 C pure rutile phase was formed. The Mössbauer spectra for all temperatures were similar, the luminescence intensity decreases with increasing calcination temperature.
However, Ningthoujam et al. [106] reported on a TiO 2 :Eu anatase phase, which showed two Mössbauer absorption bands, corresponding to Eu 3 + and Eu 2 + with isomer shifts of 0.62 mm/s and 13.04 mm/s, respectively. The samples were annealed at 500 C and 900 C and the spectra showed one peak corresponding to Eu 3 + with isomer shifts of 0.48 mm/s and 0.64 mm/s, respectively. In the 900 C annealed sample, the Eu 2 Ti 2 O 7 phase was formed, which did not show any luminescence. Annealing of the sample at temperatures from 300 C to 1000 C did not change the isomer shift significantly.
Ti 2 O 7 (cubic) pyrochlores fluoresce below room temperature [87]. The Mössbauer spectra show Eu 3 + with a large quadrupole interaction of 19 mm/s and a pronounced anisotropic f-factor (Goldanski-Karyagin Effect). The isomer shift is 0.8 mm/s.

4.8. Nitrides

Europium nitrodosilicate Eu 2 SiN 3 has mixed valence and its Mössbauer spectrum contains lines from both Eu 2 + and Eu 3 + [50]. For GaN:Eu, Mössbauer spectra showed that the Eu was preferentially in the trivalent state [135]. EuSi 2 O 2 N 2 showed an emission in the yellow spectral range and the Mössbauer spectra showed only Eu 2 + at 78 K [52].

4.9. Sulfides

CaS has a rock-salt structure and Eu 2 + is supposed to replace Ca 2 + in the lattice. For CaS:Eu luminophors, Danilkin et al. [136] observed only one Mössbauer absorption peak, corresponding to Eu 3 + ions. Pham-Thi et al. [103] synthesized CaS:Eu using the flux method with Na and K. The Mössbauer spectrum of the sample prepared with Na polysulphide flux showed only Eu 3 + absorption while with K flux both ions were detected.

4.10. ZrO2

Mössbauer spectra of ZrO 2 :Eu doped with 1 and 2 mol % Eu 2 O 3 showed only Eu 3 + with an isomer shift independent of Eu concentrations, while the linewidth decreases with increasing Eu concentration indicating that the europium environment is changing [127]. It was concluded that, at low concentrations, europium ions occupy sites in both the tetragonal and monoclinic structure and with increasing Eu concentration the incorporated ions are mainly substituted into the tetragonal structure substituting Zr 4 + .

4.11. Yttrium Aluminum Garnet (YAG:Eu)

Constantinescu et al. [81] used Mössbauer spectroscopy to investigate structural changes during the phase transition from amorphous to crystalline in yttrium aluminum garnet (Y 3 Al 5 O 12 ). Annealing the YAG with temperatures lower than the phase transition temperature, which is between 900 C and 915 C [137], revealed two peaks in the Mössbauer spectra, while annealing with higher temperatures gave one transmission peak. For annealing at temperatures from 930 C to 1400 C, an increase in crystalline size was obtained from X-ray diffraction measurements and a decrease in photoluminescence linewidths resulting from the higher crystallinity as well as a slight increase in Mössbauer absorption area was observed for all investigated temperatures [81].

4.12. Discussion

4.12.1. Correlation of Isomer Shift with Bond Length and Covalency

An increase in isomer shift corresponds to an increase in covalency (see Table 1). It is seen that for luminescent materials there is a correlation between the isomer shift and the bond length Eu–X (Figure 12), since an increase in the bond length decreases the local density at the europium sites and hence the electron density at the Eu nuclei. Such a correlation has been found for Eu 3 + in Y 2 x Eu x O 3 Hintzen (Figure 11) [66,119], Eu 3 + in Ln 2 O 3 compounds [66,86], Y 2 O 3 :Eu 3 + nanoparticles and Y 2 O 3 @Eu 3 + core-shell structures [67,116], Eu 2 + in oxides, sulphides and selenides [138].
Related correlations also exist between the isomer shift and europium concentration, which result from the difference in size between the europium and host ions e.g., Eu 2 + and Eu 3 + in La 1 x Eu x MgAl 11 O 19 [99] and Eu 2 + fluorogermanate glasses (Nemov et al. [126].)
Glass correlations are found between the isomer shift and the electronegatively (or size) of the modifying cations in sodium silicate M 3 Na 8 Si 13 O 33 and aluminophosphate M 5 Al 4 P 25 O 76 glasses (M = Mg, Ca, Sr), as shown in Figure 9 [112].

4.12.2. Determination of Site Occupancies

The occupancy of each site when there are several sites in a material may be determined since there are small differences between the isomer shift of them by computer fitting the spectrum. The unresolved and overlapping contributions from each site leads to a broadened spectrum.
The lanthanide sesquioxides with the general formula Ln 2 O 3 have complicated crystal structures. In the cubic system, there are two different sites for the lanthanide ions: 25% in a more symmetric site C 3 i (S 6 ) and 75% in a less symmetric site C 2 , which shows quadrupole broadening [66]. For Sc 2 O 3 :Eu and In 2 O 3 :Eu the two sites could be resolved [66,86]. Their average isomer shifts are 1.27 mm/s and 1.38 mm/s respectively. For Sc 1.9 Eu 0.1 O 3 the shifts for the sites C 3 i and C 2 are 0.70 mm/s and 2.59 mm/s, respectively [66]. In other compounds, the difference in shift is less and is determined from the (slightly asymmetrical) line broadening. Hintzen et al. [66] showed that the difference in isomer shift between the C 3 i and C 2 sites increases linearly with decreasing lattice parameter a. The decrease of the Eu–O distance increases the electron density, which affects the C 3 i site more than the C 2 site, as shown in Figure 10 [66,117]. Concas et al. [86,119] obtained in cubic nanocrystalline Y 2 O 3 an occupational probability of the C 3 i and C 2 sites of 27% and 73%, respectively compared with bulk material values of 22% ( C 3 i ) and 78% ( C 2 ) [119]. Concas et al. [118] also investigated Lu 1.8 Eu 0.2 O 3 . They prepared the samples in four different ways: (1) combustion with urea; (2) combustion with urea and sintered; (3) combustion with glycine; and (4) combustion with glycine and sintered. All samples show an isomer shift of 1.25 mm/s averaged over the two different sites C 3 i and C 2 . The samples show a variation in site occupancy with the different preparation methods. For the nanocrystalline powders (1) and (3), approximately 20–23% of the Eu ions are in C 3 i symmetry, while for the ceramics only 15–16% of the Eu ions are in C 3 i symmetry. Thus, there is preferential occupation of the C 2 site, especially for the (spherical) ceramic samples, which causes a higher fluorescence intensity since the 5 D 0 to 7 F 2 transition is forbidden. In monoclinic Gd 2 O 3 , the Eu 3 + ions are equally distributed over the three different crystallographic sites [66].
In Y 2 x Eu x WO 6 :Eu, the Eu 3 + ions occupy three different sites, two with C 2 and one with C 1 symmetry. Small differences in covalency produce the isomer shifts obtained of 1.25 mm/s, 0.65 mm/s, and 0.05 mm/s, respectively and which increased with Eu 3 + concentration [123].
BaCa 2 Y 6 O 12 :Eu 3 + has two different sites with isomer shifts of 0.2 and 1.5 mm/s. At the site with the lower shift, Eu 3 + replaces Y 3 + and for the higher Eu 2 + replaces Ca 2 + . The Y 3 + site is preferred with respect to the Ca 2 + site by about a factor of 2 [53].

5. Conclusions

We have reviewed the use of Mössbauer spectroscopy for characterizing luminescent materials activated by europium. Mössbauer spectroscopy is a powerful probe of europium as it is element-specific and can provide knowledge of the valence state, covalency, site symmetry, and occupation and coordination number, all of which may be important for the study and development of luminescent materials. The large difference in isomer shift between Eu 2 + and Eu 3 + enables the ionic state to be identified and the relative amounts of each to be determined in a host material. The spectra confirm that europium usually substitutes as Eu 2 + for a divalent ion like Ca 2 + and as Eu 3 + for a trivalent ion like Al 3 + or Y 3 + , but that both oxidation states may sometimes occur together since Eu 2 + easily oxidizes to Eu 3 + . The shift also depends upon the ligand X since it affects the electron density at the europium nuclei. Roughly, it scales with the local density, and so increases with increasing coordination number and decreasing Eu–X bond length or ligand diameters. This means it increases for strong covalent bonding (or low electronegativity) and for higher europium concentrations. The quadrupole splitting gives information about local symmetry and can be important for identifying or confirming the crystal site. Mössbauer spectra can also distinguish between crystallographic sites with the same ionic state and measure site occupancy in unusual circumstances, for example, in nanoparticles. Photoluminescence spectroscopy can clearly distinguish that Eu 2 + and Eu 3 + : Eu 2 + are characterized by a broad emission, while Eu 3 + shows narrow emission lines, but, as the spectra overlap, it cannot give a quantitative estimate of the relative amounts of each. In summary, Mössbauer spectroscopy is a unique and leading technique in the characterization of materials containing multi-valent elements with distinctly different isomer shifts.

Acknowledgments

The authors wish to thank the “Ministerium für Innovation, Wissenschaft und Forschung des Landes Nordrhein-Westfalen” for its financial support to the Fraunhofer Application Center for Inorganic Phosphors in Soest.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Blasse, G.; Grabmaier, B.C. Luminescent Materials; Springer: Berlin/Heidelberg, Germany, 1994. [Google Scholar]
  2. Zhang, Y.; Xu, J.; Cui, Q.; Yang, B. Eu3+-doped Bi4Si3O12 red phosphor for solid state lighting: Microwave synthesis, characterization, photoluminescence properties and thermal quenching mechanisms. Sci. Rep. 2017, 7, 42464. [Google Scholar] [CrossRef] [PubMed]
  3. Pradal, N.; Potdevin, A.; Chadeyron, G.; Bonville, P.; Caillier, B.; Mahiou, R. Spectroscopic study and enhanced thermostability of combustion-derived BaMgAl10O17:Eu2+ blue phosphors for solid-state lighting. Opt. Mater. 2017, 64, 334–344. [Google Scholar] [CrossRef]
  4. Suyver, J.F.; Meijerink, A. Europium safeguards the euro. Chemisch2Weekblad 2002, 98, 12–13. [Google Scholar]
  5. Furniss, D.; Harris, E.A.; Hollis, D.B. EPR of Gd3+ and Eu2+ in fluorozirconate glasses. J. Phys. C Solid State Phys. 1987, 20, L147–L150. [Google Scholar] [CrossRef]
  6. MacFarlane, D.R.; Newman, P.J.; Cashion, J.D.; Edgar, A. In situ generation of Eu2+ in glass-forming melts. J. Non-Cryst. Solids 1999, 256–257, 53–58. [Google Scholar] [CrossRef]
  7. Korthout, K.; Van den Eeckhout, K.; Botterman, J.; Nikitenko, S.; Poelman, D.; Smet, P.F. Luminescence and X-ray absorption measurements of persistent SrAl2O4:Eu,Dy powders: Evidence for valence state changes. Phys. Rev. B 2011, 84, 085140. [Google Scholar] [CrossRef]
  8. Korthout, K.; Parmentier, A.B.; Smet, P.F.; Poelman, D. A XAS study of the luminescent Eu centers in thiosilicate phosphors. Phys. Chem. Chem. Phys. 2013, 15, 8678–8683. [Google Scholar] [CrossRef] [PubMed]
  9. Mössbauer, R.L. Kernresonanzfluoreszenz von Gammastrahlung in Ir191. Z. Phys. 1958, 151, 124–143. [Google Scholar] [CrossRef]
  10. Kalvius, M.; Kienle, P. (Eds.) The Rudolf Mössbauer Story—His Scientific Work and Its Impact on Science and History; Springer: Berlin/Heidelberg, Germany, 2012. [Google Scholar]
  11. Grandjean, F.; Long, G.J. Mössbauer Spectroscopy Applied to Inorganic Chemistry. In Modern Inorganic Chemistry; Springer: Boston, MA, USA, 1989; Chapter Mössbauer Spectroscopy of Europium-Containing Compounds; pp. 513–597. [Google Scholar]
  12. Mössbauer Effect Data Center. Available online: http://www.medc.dicp.ac.cn (accessed on 5 March 2018).
  13. Gonser, U. Application of the Mössbauer effect in materials science. Mater. Sci. Eng. 1968, 3, 1–23. [Google Scholar] [CrossRef]
  14. Woude, F.V.D.; Sawatzky, G. Mössbauer effect in iron and dilute iron based alloys. Phys. Rep. 1974, 12, 335–374. [Google Scholar] [CrossRef]
  15. Kurkjian, C. Mössbauer spectroscopy in inorganic glasses. J. Non-Cryst. Solids 1970, 3, 157–194. [Google Scholar] [CrossRef]
  16. Müller-Warmuth, W.; Eckert, H. Nuclear magnetic-resonance and Mössbauer spectroscopy of glasses. Phys. Rep. 1982, 88, 91–149. [Google Scholar] [CrossRef]
  17. Johnson, J.A.; Johnson, C.E. Mössbauer spectroscopy as a probe of silicate glasses. J. Phys. Condens. Matter 2005, 17, R381. [Google Scholar] [CrossRef]
  18. Pandelia, M.E.; Lanz, N.D.; Booker, S.J.; Krebs, C. Mössbauer spectroscopy of Fe/S proteins. Biochim. Biophys. Acta (BBA)-Mol. Cell Res. 2015, 1853, 1395–1405. [Google Scholar] [CrossRef] [PubMed]
  19. Boncher, W.; Dalafu, H.; Rosa, N.; Stoll, S. Europium chalcogenide magnetic semiconductor nanostructures. Coord. Chem. Rev. 2015, 289–290, 279–288. [Google Scholar] [CrossRef]
  20. Dieke, G.H.; Crosswhite, H.M. The Spectra of the Doubly and Triply Ionized Rare Earths. Appl. Opt. 1963, 2, 675–686. [Google Scholar] [CrossRef]
  21. Martin, W.C.; Zalubas, R.; Hagan, L. Atomic Energy Levels—The Rare-Earth Elements: The Spectra of Lanthanum, Cerium, Praseodymium, Neodymium, Promethium, Samarium, Europium, Gadolinium, Terbium, Dysprosium, Holmium, Erbium, Thulium, Ytterbium, and Lutetitum; U.S. Department of Commerce, National Bureau of Standards: Washington, DC, USA, 1978.
  22. Liu, G. Advances in the theoretical understanding of photon upconversion in rare-earth activated nanophosphors. Chem. Soc. Rev. 2015, 44, 1635–1652. [Google Scholar] [CrossRef] [PubMed]
  23. Poulain, M.; Poulain, M.; Lucas, J.; Brun, P. Verres fluores au tetrafluorure de zirconium proprietes optiques d’un verre dope au Nd3+. Mater. Res. Bull. 1975, 10, 243. [Google Scholar] [CrossRef]
  24. Digonnet, M.J.F. (Ed.) Rare-Earth-Doped Fiber Lasers and Amplifiers, Revised and Expanded; Marcel Dekker, Inc.: New York, NY, USA, 2001. [Google Scholar]
  25. Zhu, X.; Peyghambarian, N. High-Power ZBLAN Glass Fiber Lasers: Review and Prospect. Adv. OptoElectron. 2010, 2010, 1–23. [Google Scholar] [CrossRef]
  26. Calandrelli, E. Made in Space Plans to Create a Superior Optical Fiber in Microgravity. Available online: https://techcrunch.com/2016/07/13/made-in-space-plans-to-create-a-superior-optical-fiber-in-microgravity/ (accessed on 4 January 2018).
  27. Schweizer, S.; Hobbs, L.W.; Secu, M.; Spaeth, J.M.; Edgar, A.; Williams, G.V.M. Photostimulated luminescence in Eu-doped fluorochlorozirconate glass ceramics. Appl. Phys. Lett. 2003, 83, 449. [Google Scholar] [CrossRef]
  28. Johnson, C.E.; Vu, M.; Johnson, J.A.; Brown, D.E.; Weber, J.K.R.; Paßlick, C.; Schweizer, S. Mössbauer spectroscopy of europium-containing glasses: optical activator study for X-ray image plates. Hyperfine Interact. 2014, 226, 797–801. [Google Scholar] [CrossRef] [PubMed]
  29. Downing, E.; Hesselink, L.; Ralston, J.; Macfarlane, R. A three-color, solid-state, three-dimensional display. Science 1996, 273, 1185–1189. [Google Scholar] [CrossRef]
  30. Trupke, T.; Green, M.A.; Würfel, P. Improving solar cell efficiencies by up-conversion of sub-band-gab light. J. Appl. Phys. 2002, 92, 4117. [Google Scholar] [CrossRef]
  31. Ahrens, B.; Löper, P.; Goldschmidt, J.C.; Glunz, S.; Henke, B.; Miclea, P.T.; Schweizer, S. Neodymium-doped fluorochlorozirconate glasses as an upconversion model system for high efficiency solar cells. Phys. Status Solidi A 2008, 205, 2822–2830. [Google Scholar] [CrossRef]
  32. Johnson, J.A.; Schweizer, S.; Henke, B.; Chen, G.; Woodford, J.; Newman, P.J.; MacFarlane, D.R. Eu-activated fluorochlorozirconate glass-ceramic scintillators. J. Appl. Phys. 2006, 100, 034701. [Google Scholar] [CrossRef]
  33. Johnson, J.A.; Schweizer, S.; Lubinsky, A.R. A Glass-Ceramic Plate for Mammography. J. Am. Ceram. Soc. 2007, 90, 693–698. [Google Scholar] [CrossRef]
  34. Schweizer, S.; Hobbs, L.W.; Secu, M.; Spaeth, J.M.; Edgar, A.; Williams, G.V.M.; Hamlin, J. Photostimulated luminescence from fluorochlorozirconate glass ceramics and the effect of crystallite size. J. Appl. Phys. 2005, 97. [Google Scholar] [CrossRef]
  35. Shafer, M.W.; Suits, J.C. Preparation and Faraday Rotation of Divalent Europium Glasses. J. Am. Ceram. Soc. 1966, 49, 261–264. [Google Scholar] [CrossRef]
  36. Fujita, K.; Tanaka, K.; Hirao, K.; Soga, N. Mössbauer Spectroscopy of Borate Glasses Containing Divalent Europium Ions. J. Am. Ceram. Soc. 1998, 81, 1845–1851. [Google Scholar] [CrossRef]
  37. Hangleiter, T.; Koschnick, F.K.; Spaeth, J.M.; Nuttall, R.H.D.; Eachus, R.S. Temperature dependence of the photostimulated luminescence of X-irradiate BaFBr:Eu2+. J. Phys. Condens. Matter 1990, 2, 6837. [Google Scholar] [CrossRef]
  38. Li, Y.; Gecevicius, M.; Qiu, J. Long persistent phosphors - from fundamentals to applications. Chem. Soc. Rev. 2016, 45, 2090–2136. [Google Scholar] [CrossRef] [PubMed]
  39. Rojas-Hernandez, R.E.; Rubio-Marcos, F.; Rodriguez, M.A.; Fernandez, J.F. Long lasting phosphors: SrAl2O4:Eu, Dy as the most studied material. Renew. Sustain. Energy Rev. 2018, 81, 2759–2770. [Google Scholar] [CrossRef]
  40. Kitsinelis, S.; Kitsinelis, S. Light Sources, Second Edition: Basics of Lighting Technologies and Applications; CRC Press: Boca Raton, FL, USA, 2015. [Google Scholar]
  41. Liu, R.S. (Ed.) Phosphors, up Conversion Nano Particles, Quantum Dots and Their Applications; Springer: Berlin/Heidelberg, Germany, 2017; Volume 1. [Google Scholar]
  42. Hölsä, J.; Aitasalo, T.; Jungner, H.; Lastusaari, M.; Niittykoski, J.; Spano, G. Role of defect states in persistent luminescence materials. J. Alloys Compd. 2004, 374, 56–59. [Google Scholar] [CrossRef]
  43. Aitasalo, T.; Dereń, P.; Hölsä, J.; Jungner, H.; Krupa, J.C.; Lastusaari, M.; Legendziewicz, J.; Niittykoski, J.; Strȩk, W. Persistent luminescence phenomena in materials doped with rare earth ions. J. Solid State Chem. 2003, 171, 114–122. [Google Scholar] [CrossRef]
  44. Yuan, Z.X.; Chang, C.K.; Mao, D.L.; Ying, W. Effect of composition on the luminescent properties of Sr4Al14O25: Eu2+, Dy3+ phosphors. J. Alloys Compd. 2004, 377, 268–271. [Google Scholar] [CrossRef]
  45. Clabau, F.; Garcia, A.; Bonville, P.; Gonbeau, D.; Mercier, T.L.; Deniard, P.; Jobic, S. Fluorescence and phosphorescence properties of the low temperature forms of the MAl2Si2O8:Eu2+ (M = Ca, Sr, Ba) compounds. J. Solid State Chem. 2008, 181, 1456–1461. [Google Scholar] [CrossRef]
  46. Pathak, P.K.; Kurchania, R. Synthesis and thermoluminescence properties of SrAl2O4 (EU) phosphor irradiated with cobalt-60, 6 MV and 16 MV photon beams. Radiat. Phys. Chem. 2015, 117, 48–53. [Google Scholar] [CrossRef]
  47. Sweet, M.; Rennie, J. Thermoluminescence dosimetry using doped calcium sulphide. J. Cryst. Growth 1988, 86, 942–947. [Google Scholar] [CrossRef]
  48. Title, R.S. Paramagnetic Resonance Detection of the Optical Excitation of an Infrared Stimulable Phosphor. Phys. Rev. Lett. 1959, 3, 273–274. [Google Scholar] [CrossRef]
  49. Riedel, R.; Chen, I.W. (Eds.) Ceramics Science and Technology, Volume 1: Structures; Wiley-VCH Verlag GmbH & Co. KGaA: Weinheim, Germany, 2008. [Google Scholar]
  50. Zeuner, M.; Pagano, S.; Matthes, P.; Bichler, D.; Johrendt, D.; Harmening, T.; Pöttgen, R.; Schnick, W. Mixed Valence Europium Nitridosilicate Eu2SiN3. J. Am. Chem. Soc. 2009, 131, 11242–11248. [Google Scholar] [CrossRef] [PubMed]
  51. ten Kate, O.M.; Vranken, T.; van der Kolk, E.; Jansen, A.P.; Hintzen, H.T. Optical properties of Eu2+/Eu3+ mixed valence, silicon nitride based materials. J. Solid State Chem. 2014, 213, 126–131. [Google Scholar] [CrossRef]
  52. Stadler, F.; Oeckler, O.; Höppe, H.A.; Möller, M.H.; Pöttgen, R.; Mosel, B.D.; Schmidt, P.; Duppel, V.; Simon, A.; Schnick, W. Crystal Structure, Physical Properties and HRTEM Investigation of the New Oxonitridosilicate EuSi2O2N2. Chem. Eur. J. 2006, 12, 6984–6990. [Google Scholar] [CrossRef] [PubMed]
  53. Hintzen, H.T.; Van Hal, H.A.M.; Langereis, C.; Denissen, C.J.M. Crystal chemistry of BaCa2Ln6O12 (Ln = In, Sc, Y, Dy, Ho, Er, Tm, Yb, Lu) and luminescence of europium- and terbium-activated BaCa2Y6O12. J. Less Common Met. 1989, 155, 291–305. [Google Scholar] [CrossRef]
  54. Zych, E.; Meijerink, A.; de Mello Donegá, C. Quantum efficiency of europium emission from nanocrystalline powders of Lu2O3:Eu. J. Phys. Condens. Matter 2003, 15, 5145. [Google Scholar] [CrossRef]
  55. Meng, Q.; Chen, B.; Xu, W.; Wang, X.; Xie, Y.; Lu, S.; Lai, H.; Yang, Y.; Zhao, X.; Di, W.; et al. Fluorescence lifetime and quantum efficiency of Y2O3:Eu nanocrystals. In Proceedings of the ICO20: Display Devices and Systems, Changchun, China, 21–26 August 2005; SPIE: Bellingham, WA, USA, 2016; Volume 6030, pp. 60300I-1–60300I-6. [Google Scholar] [CrossRef]
  56. Steudel, F.; Loos, S.; Ahrens, B.; Schweizer, S. Quantum efficiency and energy transfer processes in rare-earth doped borate glass for solid-state lighting. J. Lumin. 2016, 170, 770–777. [Google Scholar] [CrossRef]
  57. Loos, S.; Steudel, F.; Ahrens, B.; Leonard, R.L.; Johnson, J.A.; Schweizer, S. Temperature-dependent luminescence of Tb3+ and Eu3+ single-doped glasses for LED applications. Phys. Status Solidi C 2015, 12, 1359–1364. [Google Scholar] [CrossRef]
  58. Wetenkamp, L.; West, G.; Többen, H. Co-doping effects in erbium3+- and holmium3+-doped ZBLAN glasses. J. Non-Cryst. Solids 1992, 140, 25–30. [Google Scholar] [CrossRef]
  59. Tohmon, G.; Sato, H.; Ohya, J.; Fujita, T. Energy transfer in Tm:Eu codoped fluorozirconate fiber. J. Appl. Phys. 1993, 73, 1528–1530. [Google Scholar] [CrossRef]
  60. Steudel, F.; Loos, S.; Ahrens, B.; Schweizer, S. Luminescent borate glass for efficiency enhancement of CdTe solar cells. J. Lumin. 2015, 164, 76–80. [Google Scholar] [CrossRef]
  61. Song, P.; Zhang, C.; Zhu, P. Efficiency stokes shifting luminescent materials: Ce3+-Tb3+-Eu3+ toward solar energy applications. Ceram. Int. 2016, 42, 11417–11421. [Google Scholar] [CrossRef]
  62. Meza-Rocha, A.; Lozada-Morales, R.; Speghini, A.; Bettinelli, M.; Caldiño, U. White light generation in Tb3+/Eu3+/Dy3+ triply-doped Zn(PO3)2 glass. Opt. Mater. 2016, 51, 128–132. [Google Scholar] [CrossRef]
  63. Van Dijk, J.M.F.; Schuurmans, M.F.H. On the nonradiative and radiative decay rates and a modified exponential energy gap law for 4f–4f transitions in rare earth ions. J. Chem. Phys. 1983, 78, 5317. [Google Scholar] [CrossRef]
  64. Shojiya, M.; Takahashi, M.; Kanno, R.; Kawamoto, Y.; Kadono, K. Optical transitions of Er3+ ions in ZnCl2-based glass. J. Appl. Phys. 1997, 82, 6259–6266. [Google Scholar] [CrossRef]
  65. Bendow, B.; Banerjee, P.K.; Drexhage, M.G.; Goltman, J.; Mitra, S.S.; Moynihan, C.T. Comparative study of vibrational characteristics of fluorozirconate and fluorohafnate glasses. J. Am. Ceram. Soc. 1982, 65, C8–C9. [Google Scholar] [CrossRef]
  66. Hintzen, H.; van Noort, H. Investigation of luminescent Eu-doped sesquioxides Ln2O3 (Ln = In, Sc, Y, La, Gd, Lu) and some mixed oxides by 151Eu Mössbauer spectroscopy. J. Phys. Chem. Solids 1988, 49, 873–881. [Google Scholar] [CrossRef]
  67. Bohus, G.; Hornok, V.; Oszkó, A.; Vértes, A.; Kuzmann, E.; Dékány, I. Structural and luminescence properties of Y2O3:Eu3+ core-shell nanoparticles. Colloids Surfaces A Physicochem. Eng. Aspects 2012, 405, 6–13. [Google Scholar] [CrossRef]
  68. Bartram, R.; Lempicki, A.; Kappers, L.; Hamilton, D. Hole traps in Lu2O3:Eu ceramic scintillators. II. Radioluminescence and thermoluminescence. J. Lumin. 2004, 106, 169–176. [Google Scholar] [CrossRef]
  69. Zych, E.; Trojan-Piegza, J.; Kępiński, L. Homogeneously precipitated Lu2O3:Eu nanocrystalline phosphor for X-ray detection. Sens. Actuators B Chem. 2005, 109, 112–118. [Google Scholar] [CrossRef]
  70. Morkoç, H. Handbook of Nitride Semiconductors and Devices: GaN-Based Optical and Electronic Devices; Wiley-VCH Verlag GmbH & Co. KGaA: Weinheim, Germany, 2009; Volume 3. [Google Scholar] [CrossRef]
  71. Hansel, R.A.; Desai, S.K.; Allison, S.W.; Heyes, A.L.; Walker, D.G. Emission lifetimes of europium-doped pyrochlores for phosphor thermometry. J. Appl. Phys. 2010, 107, 016101. [Google Scholar] [CrossRef]
  72. Luo, M.; Cheng, K.; Weng, W.; Song, C.; Du, P.; Shen, G.; Xu, G.; Han, G. Enhanced Luminescence of Eu-Doped TiO2 Nanodots. Nanoscale Res. Lett. 2009, 4, 809–813. [Google Scholar] [CrossRef] [PubMed]
  73. Frindell, K.L.; Bartl, M.H.; Popitsch, A.; Stucky, G.D. Sensitized Luminescence of Trivalent Europium by Three-Dimensionally Arranged Anatase Nanocrystals in Mesostructured Titania Thin Films. Angew. Chem. Int. Ed. 2002, 41, 959–962. [Google Scholar] [CrossRef]
  74. Chen, X.; Mao, S.S. Titanium Dioxide Nanomaterials: Synthesis, Properties, Modifications, and Applications. Chem. Rev. 2007, 107, 2891–2959. [Google Scholar] [CrossRef] [PubMed]
  75. Shionoya, S.; Yen, W.M.; Yamamoto, H. (Eds.) Phosphor Handbook; CRC Press: Boca Raton, FL, USA, 2006. [Google Scholar]
  76. Shirmane, L.; Feldmann, C.; Pankratov, V. Comparing the luminescence processes of YVO4:Eu and core-shell YVO4@YF3 nanocrystals with bulk-YVO4:Eu. Phys. B Condens. Matter 2017, 504, 80–85. [Google Scholar] [CrossRef]
  77. Giaume, D.; Buissette, V.; Lahlil, K.; Gacoin, T.; Boilot, J.P.; Casanova, D.; Beaurepaire, E.; Sauviat, M.P.; Mercuri, A.; Alexandrou, A. New biological labels based on functionalized YVO4:Eu nanoparticles. MRS Online Proc. Libr. Arch. 2004, 845, AA6.2. [Google Scholar] [CrossRef]
  78. Zhao, Y.; Shao, M.; Liu, S.; Zhang, Z.; Lin, H. Hydrothermal synthesis of lanthanide orthovanadate: EuVO4 particles and their fluorescence application. CrystEngComm 2012, 14, 8033–8036. [Google Scholar] [CrossRef]
  79. Smits, K.; Grigorjeva, L.; Millers, D.; Sarakovskis, A.; Opalinska, A.; Fidelus, J.D.; Lojkowski, W. Europium doped zirconia luminescence. Opt. Mater. 2010, 32, 827–831. [Google Scholar] [CrossRef]
  80. Tamrakar, R.K.; Bisen, D.P.; Upadhyay, K. Photoluminescence behavior of ZrO2:Eu3+ with variable concentration of Eu3+ doped phosphor. J. Radiat. Res. Appl. Sci. 2015, 8, 11–16. [Google Scholar] [CrossRef]
  81. Constantinescu, S.; Georgescu, S.; Bibicu, I.; Chinie, A.A.; Stefan, A.S.; Toma, O. Mossbauer and optical investigation of Eu:YAG nanocrystals synthesized by a sol-gel method. Rom. J. Phys. 2007, 52, 295–307. [Google Scholar]
  82. Yang, L.; Lv, Z.; Chen, M.; Liu, S. Combination of translucent Eu:YAG glass ceramic with LED chip. In Proceedings of the 2012 IEEE 62nd Electronic Components and Technology Conference, San Diego, CA, USA, 29 May–1 June 2012; pp. 2145–2149. [Google Scholar] [CrossRef]
  83. Mössbauer Effect Data Center. 151Eu Isotope Properties. Available online: http://www.medc.dicp.ac.cn/Resources-isotopes/Resource-Eu.php (accessed on 5 March 2018).
  84. Bibicu, I.; Cretu, C. Mössbauer backscattering measurements on Eu-151. Rom. J. Phys. 2009, 54, 515–519. [Google Scholar]
  85. Bibicu, I. Some comments on 151Eu Mössbauer spectroscopy. Eur. Phys. J. Appl. Phys. 2013, 62, 11302. [Google Scholar] [CrossRef]
  86. Concas, G.; Congiu, F.; Spano, G.; Bettinelli, M.; Speghini, A. Mössbauer Investigation of Eu3+ Site Occupancy and Eu–O Covalency in Y2O3 and Gd2O3 Nanocrystals. Z. Naturforsch. 2001, 56, 267–272. [Google Scholar] [CrossRef]
  87. Chien, C.L.; Sleight, A.W. Mössbauer effect studies of europium pyrochlores. Phys. Rev. B 1978, 18, 2031–2038. [Google Scholar] [CrossRef]
  88. Sanchez, J.P.; Friedt, J.M.; Westerholt, K.; Bach, H. 151Eu Mössbauer study of the EuxLa1−xS solid solution. Phys. Rev. B 1986, 33, 4514–4521. [Google Scholar] [CrossRef]
  89. Kuzmann, E.; Homonnay, Z.; Nagy, S.; Nomura, K. Handbook of Nuclear Chemistry; Springer Science & Business Media: Boston, MA, USA, 2011; Chapter Mössbauer Spectroscopy; pp. 1379–1446. [Google Scholar] [CrossRef]
  90. Concas, G.; Congiu, F.; Spano, G.; Speghini, A.; Gatterer, K. Mössbauer investigation of rare earth sites in europium containing glasses. J. Non-Cryst. Solids 1998, 232–234, 341–345. [Google Scholar] [CrossRef]
  91. Weber, J.K.R.; Vu, M.; Paßlick, C.; Schweizer, S.; Brown, D.E.; Johnson, C.E.; Johnson, J.A. The oxidation state of europium in halide glasses. J. Phys. Condens. Matter 2011, 23, 495402. [Google Scholar] [CrossRef] [PubMed]
  92. Selling, J.; Bielemeier, B.; Wortmann, G.; Johnson, J.A.; Alp, E.E.; Chen, T.; Brown, D.E.; Johnson, C.E.; Schweizer, S. Paramagnetic hyperfine splitting in the 151Eu Mössbauer spectra of CaF2:Eu2+. Phys. Rev. B 2008, 77, 224442. [Google Scholar] [CrossRef] [PubMed]
  93. Coey, J.M.D.; McEvoy, A.; Shafer, M.W. Mössbauer study of europium in fluorozirconate glass. J. Non-Cryst. Solids 1981, 43, 387–392. [Google Scholar] [CrossRef]
  94. Pfau, C.; Paßlick, C.; Gray, S.K.; Johnson, J.A.; Johnson, C.E.; Schweizer, S. Mössbauer spectroscopy of europium-doped fluorochlorozirconate glasses and glass ceramics: optimization of storage phosphors in computed radiography. J. Phys. Condens. Matter 2013, 25, 205402. [Google Scholar] [CrossRef] [PubMed]
  95. Arakawa, T.; Nagata, N. Luminescence properties of A1−xEuxAl12O19S6 (A=Ca, Sr, Ba). J. Alloys Compd. 2006, 408–412, 864–866. [Google Scholar] [CrossRef]
  96. Arakawa, T.; Fuchigami, H. Preparation and Europium-151 Mössbauer Spectroscopy of Ca1−xEuxAl12O19. Hyperfine Interactions (C); Thomas, M.F., Williams, J.M., Gibb, T.C., Eds.; Springer: Dordrecht, The Netherlands, 2002; pp. 359–361. [Google Scholar] [CrossRef]
  97. Tezuka, K.; Tokuhara, Y.; Wakeshima, M.; Shan, Y.J.; Imoto, H.; Hinatsu, Y. Crystal Structures and Properties of Europium Aluminum Oxynitride Eu2AlO3.75N0.1 and Europium Aluminum Oxide EuAl2O4. Inorg. Chem. 2013, 52, 12972–12979. [Google Scholar] [CrossRef] [PubMed]
  98. Boolchand, P.; Mishra, K.C.; Raukas, M.; Ellens, A.; Schmidt, P.C. Occupancy and site distribution of europium in barium magnesium aluminate by 151Eu Mössbauer spectroscopy. Phys. Rev. B 2002, 66, 134429. [Google Scholar] [CrossRef]
  99. Tronc, E.; Saber, D.; Lejus, A.M.; Vivien, D. 151Eu Mössbauer investigation of La1−xEuxMgAl11O19. J. Less Common Met. 1985, 111, 321–327. [Google Scholar] [CrossRef]
  100. Fraknóy-Körös, V.; Gelencsér, P.; Nagy, I.C.; Vértes, A. Application of Mössbauer spectroscopy to investigate different europium(II) and (III)-activated phosphors. Radiochem. Radioanal. Lett. 1980, 44, 337–346. [Google Scholar]
  101. Hintzen, H.T.; Denissen, C.J.M.; van Noort, H.M. 151Eu Mössbauer spectroscopic study of the phosphor SrAl12O19:Eu with the magnetoplumbite structure. Mater. Res. Bull. 1989, 24, 247–259. [Google Scholar] [CrossRef]
  102. Winterer, M.; Mörsen, E.; Mosel, B.D.; Müller-Warmuth, W. Paramagnetic hyperfine structure in 151Eu Mössbauer spectra of Eu2+ ions in borate glasses. J. Phys. C Solid State Phys. 1987, 20, 5389–5399. [Google Scholar] [CrossRef]
  103. Pham-Thi, M.; Ruelle, N.; Tronc, E.; Simons, D.; Vivien, D. Electron Spin Resonance and Mössbauer Studies of CaS:Eu Synthesized Using the Flux Method. Jpn. J. Appl. Phys. 1994, 33, 1876. [Google Scholar] [CrossRef]
  104. Johnson, C.E.; Costa, L.; Johnson, J.A.; Brown, D.E.; Somarajan, S.; He, W.; Dickerson, J.H. Mössbauer spectra and superparamagnetism of europium sulfide nanoparticles. J. Phys. D Appl. Phys. 2014, 47, 075001. [Google Scholar] [CrossRef]
  105. Stadnik, Z.M.; Stroink, G.; Arakawa, T. Quadrupole interactions at divalent and trivalent europium sites in several europium oxides. Phys. Rev. B 1991, 22, 12552–12558. [Google Scholar] [CrossRef]
  106. Ningthoujam, R.; Sudarsan, V.; Vatsa, R.; Kadam, R.; Jagannath; Gupta, A. Photoluminescence studies on Eu doped TiO2 nanoparticles. J. Alloys Compd. 2009, 486, 864–870. [Google Scholar] [CrossRef]
  107. Tanabe, S.; Hirao, K.; Soga, N. Local structure of rare-earth ions in fluorophosphate glasses by phonon sideband and mössbauer spectroscopy. J. Non-Cryst. Solids 1992, 142, 148–154. [Google Scholar] [CrossRef]
  108. Li, J.; Shi, Y.; Gong, J.; Chen, G. Mössbauer study of amorphous Al2O3:Eu3+. J. Mater. Sci. Lett. 1997, 16, 743–744. [Google Scholar] [CrossRef]
  109. Zhang, L.; Lu, C.; Long, Y. Synthesis, characterization of an AlPO-CJ2 analogue containing heteroatomic Eu. Chem. Commun. 2002, 2064–2065. [Google Scholar] [CrossRef]
  110. Taragin, M.F.; Eisenstein, J.C. Mössbauer-Effect Study of Europium in Glass. Phys. Rev. B 1970, 2, 3490–3494. [Google Scholar] [CrossRef]
  111. Musić, S.; Bajs, Z.; Furić, K.; Mohac̆ek, V. Mössbauer and vibrational spectra of sodium borosilicate glasses containing europium or tin ions. J. Mater. Sci. Lett. 1991, 10, 889–892. [Google Scholar] [CrossRef]
  112. Tanabe, S.; Hanada, T.; Ohyagi, T.; Soga, N. Correlation between 151Eu Mössbauer isomer shift and Judd-Ofelt Ω6 parameters of Nd3+ ions in phosphate and silicate laser glasses. Phys. Rev. B 1993, 48, 10591–10594. [Google Scholar] [CrossRef]
  113. Pavel, E.; Mihailescu, I.N.; Hening, A.; Vlad, V.I.; Tugulea, L.; Diamandescu, L.; Bibicu, I.; Chipara, M. Three-dimensional memory effect in fluorescent photosensitive glass activated by europium and cerium. Opt. Lett. 1998, 23, 1304–1306. [Google Scholar] [CrossRef] [PubMed]
  114. Tanabe, S.; Hirao, K.; Soga, N. Mössbauer spectroscopy of 151Eu in oxide crystals and glasses. J. Non-Cryst. Solids 1989, 113, 178–184. [Google Scholar] [CrossRef]
  115. Concas, G.; Congiu, F.; Muntoni, C.; Bettinelli, M.; Speghini, A. Hyperfine interactions at europium sites in oxide glasses. Phys. Rev. B 1996, 53, 6197–6202. [Google Scholar] [CrossRef]
  116. Kuzmann, E.; Vértes, A.; Bohus, G.; Hornok, V.; Oszkó, A.; Dékány, I. 151Eu Mössbauer study of luminescent Y2O3:Eu3+ core-shell nanoparticles. Hyperfine Interact. 2013, 218, 23–28. [Google Scholar] [CrossRef]
  117. Concas, G.; Spano, G.; Bettinelli, M.; Speghini, A. Distribution of Eu3+ Dopant Ions in C3i and C2 Sites of the Nanocrystalline Sc2O3:Eu Phosphor. Z. Naturforsch. 2008, 63a, 210–216. [Google Scholar] [CrossRef]
  118. Concas, G.; Spano, G.; Zych, E.; Trojan-Piegzan, J. Nano- and microcrystalline Lu2O3:Eu phosphors: variations in occupancy of C2 and S6 sites by Eu3+ ions. J. Phys. Condens. Matter 2005, 17, 2597. [Google Scholar] [CrossRef]
  119. Concas, G.; Spano, G.; Bettinellia, M.; Speghinia, A. Investigation of Eu3+ Site Occupancy in Cubic Y2O3 and Lu2O3 Nanocrystals. Z. Naturforsch. 2003, 58, 551–557. [Google Scholar] [CrossRef]
  120. Bibicu, I.; Constantinescu, S.; Diamandescu, L.; Voiculescu, A.M.; Cotoi, E. Mössbauer spectroscopy study on YVO4:Eu luminescent material. Rom. Rep. Phys. 2014, 66, 1012–1017. [Google Scholar]
  121. Li, L.P.; Li, G.S.; Xue, Y.F.; Inomatac, H. Structure, Luminescence, and Transport Properties of EuVO4. J. Electrochem. Soc. 2001, 148, J45–J49. [Google Scholar] [CrossRef]
  122. Schwung, S.; Enseling, D.; Wesemann, V.; Rytz, D.; Heying, B.; Rodewald, U.C.; Gerke, B.; Niehaus, O.; Pöttgen, R.; Jüstel, T. KYW42O8:Eu3+—A closer look on its photoluminescence and structure. J. Lumin. 2015, 159, 251–257. [Google Scholar] [CrossRef]
  123. Van Noort, H.M.; Popma, T.J.A. Concentration-dependent site occupancy in europium-doped Y2WO6 as studied by 151Eu Mössbauer spectroscopy. Solid State Commun. 1985, 55, 77–79. [Google Scholar] [CrossRef]
  124. Fujishiro, F.; Murakami, M.; Hashimoto, T.; Takahashi, M. Orange luminescence of Eu3+-doped CuLaO2 delafossite oxide. J. Ceram. Soc. Jpn. 2010, 118, 1217–1220. [Google Scholar] [CrossRef]
  125. Ovenstone, J.; Titler, P.J.; Withnall, R.; Silver, J. A Study of the Effects of Europium Doping and Calcination on the Luminescence of Titania Phosphor Materials. J. Phys. Chem. B 2001, 105, 7170–7177. [Google Scholar] [CrossRef]
  126. Nemov, S.A.; Marchenko, A.V.; Seregin, P.P. Parameters of luminescence and the local structure of Eu3+ centers in fluorogermanate glasses. Glass Phys. Chem. 2008, 34, 47–49. [Google Scholar] [CrossRef]
  127. Salas, P.; Nava, N.; Angeles-Chavez, C.; la Rosa, E.D.; Diaz-Torres, L.A. Structural and Spectroscopic Characterization of ZrO2:Eu3+ Nanoparticles. J. Nanosci. Nanotechnol. 2008, 8, 1901–1912. [Google Scholar] [CrossRef]
  128. Dhopte, S.; Muthal, P.; Kondawar, V.; Moharil, S. Luminescence in CaF2:Eu. J. Lumin. 1992, 54, 95–101. [Google Scholar] [CrossRef]
  129. Hao, Z.; Cai, J.; Li, Y. The luminescence of Eu2+ in CaF2 crystal. J. Lumin. 1988, 40–41, 393–394. [Google Scholar] [CrossRef]
  130. Mishra, K.C.; Raukas, M.; Marking, G.; Chen, P.; Boolchand, P. Investigation of Fluorescence Degradation Mechanism of Hydrated BaMgAl10O17:Eu2+ Phosphor. J. Electrochem. Soc. 2005, 152, H183–H190. [Google Scholar] [CrossRef]
  131. Fujita, K.; Tanaka, K.; Yamashita, K.; Hirao, K. Room-temperature persistent spectral hole burning of Eu3+-doped sodium borate glasses. J. Lumin. 2000, 87–89, 682–684. [Google Scholar] [CrossRef]
  132. Bauminger, E.R.; Kalvius, G.M.; Nowik, I. Mössbauer Isomer Shifts; North-Holland Publishing Company: Amsterdam, The Netherlands, 1978; Chapter 10; pp. 661–756. [Google Scholar]
  133. Large, N.; Bullock, R.; Glentworth, P.; Newton, D. Isomer shift of the Mössbauer spectrum of 151EuF3. Phys. Lett. A 1969, 29, 352–353. [Google Scholar] [CrossRef]
  134. Clifford, A.F. Mössbauer Spectroscopy of the Rare Earths. In The Mössbauer Effect and Its Application in Chemistry; American Chemical Society: Washington, DC, USA, 1967; Chapter 8; pp. 113–125. [Google Scholar] [CrossRef]
  135. Mezdrogina, M.M.; Danilovsky, E.Y.; Kuzmin, R.V. Emission sensitization and mechanisms of electron-excitation migration in structures based on III-nitrides doped with rare-earth elements (Eu, Er, Sm). Semiconductors 2010, 44, 321–328. [Google Scholar] [CrossRef]
  136. Danilkin, M.; Must, M.; Pedak, E.; Pärnoja, E.; Ryasnyi, G.; Shpinkov, I.; Shpinkova, L.; Seeman, V. Different Eu-centres in CaS:Eu, Cl. Radiat. Meas. 1995, 24, 351–354. [Google Scholar] [CrossRef]
  137. Suárez, M.; Fernández, A.; Menéndez, J.L.; Torrecillas, R. Transparent Yttrium Aluminium Garnet Obtained by Spark Plasma Sintering of Lyophilized Gels. J. Nanomater. 2009, 2009, 5. [Google Scholar] [CrossRef]
  138. Berkooz, O. Isomer shifts of 151Eu in divalent europium compounds. J. Phys. Chem. Solids 1969, 30, 1763–1767. [Google Scholar] [CrossRef]
Figure 1. Energy level diagrams of Eu 2 + (top) and Eu 3 + (bottom). The 4 f n levels are depicted as solid black lines and the closely spaced levels of 4 f n 1 5d and 4 f n 1 5p are depicted as grey bands [20,21,22].
Figure 1. Energy level diagrams of Eu 2 + (top) and Eu 3 + (bottom). The 4 f n levels are depicted as solid black lines and the closely spaced levels of 4 f n 1 5d and 4 f n 1 5p are depicted as grey bands [20,21,22].
Materials 11 00828 g001
Figure 2. Normalized PL emission spectra of Eu 2 + -doped fluorochlorozirconate glass processed at 260 C (solid curve) and 290 C (dashed curve) for 20 min. The PL was excited at 285 nm and recorded at room temperature. The 260 C processed sample only contains hexagonal BaCl 2 in a glass matrix, while the sample processed at 290 C contains mainly orthorhombic BaCl 2 in a glass matrix [34]. The quantum efficiency spectrum (on the left side) has been recorded for the 290 C processed sample.
Figure 2. Normalized PL emission spectra of Eu 2 + -doped fluorochlorozirconate glass processed at 260 C (solid curve) and 290 C (dashed curve) for 20 min. The PL was excited at 285 nm and recorded at room temperature. The 260 C processed sample only contains hexagonal BaCl 2 in a glass matrix, while the sample processed at 290 C contains mainly orthorhombic BaCl 2 in a glass matrix [34]. The quantum efficiency spectrum (on the left side) has been recorded for the 290 C processed sample.
Materials 11 00828 g002
Figure 3. Absolute PL quantum efficiency and PL emission spectra of Eu 3 + single-doped borate (black curves) and ZBLAN glass (blue curves). The Eu 3 + emissions are recorded under 395-nm excitation at room temperature [57].
Figure 3. Absolute PL quantum efficiency and PL emission spectra of Eu 3 + single-doped borate (black curves) and ZBLAN glass (blue curves). The Eu 3 + emissions are recorded under 395-nm excitation at room temperature [57].
Materials 11 00828 g003
Figure 4. Nuclear decay scheme for 151 Sm and 151 Eu [83].
Figure 4. Nuclear decay scheme for 151 Sm and 151 Eu [83].
Materials 11 00828 g004
Figure 5. Mössbauer spectra of Eu 2 Ti 2 O 7 showing quadrupole splitting [87].
Figure 5. Mössbauer spectra of Eu 2 Ti 2 O 7 showing quadrupole splitting [87].
Materials 11 00828 g005
Figure 6. Mössbauer spectra of Eu 0.65 La 0.35 S showing magnetic hyperfine splitting arising from slow electron spin relaxation [88].
Figure 6. Mössbauer spectra of Eu 0.65 La 0.35 S showing magnetic hyperfine splitting arising from slow electron spin relaxation [88].
Materials 11 00828 g006
Figure 7. Plot of isomer shifts for the Eu compounds summarized in Table 2 and Table 3.
Figure 7. Plot of isomer shifts for the Eu compounds summarized in Table 2 and Table 3.
Materials 11 00828 g007
Figure 8. Mössbauer spectra of ZBLAN samples made with (a) 5.0 mol % EuCl 2 ; and (b) 2.5 mol % EuCl 2 and 2.5 mol % EuCl 3 [91]. The solid lines represent a fit of the Mössbauer spectra using simple Lorentzian functions.
Figure 8. Mössbauer spectra of ZBLAN samples made with (a) 5.0 mol % EuCl 2 ; and (b) 2.5 mol % EuCl 2 and 2.5 mol % EuCl 3 [91]. The solid lines represent a fit of the Mössbauer spectra using simple Lorentzian functions.
Materials 11 00828 g008
Figure 9. Relationship between isomer shift and the electronegativity of network-modifying alkali (full symbols) and alkali earths (open symbols) in silicate (squares) and phosphate (circles) glasses [112].
Figure 9. Relationship between isomer shift and the electronegativity of network-modifying alkali (full symbols) and alkali earths (open symbols) in silicate (squares) and phosphate (circles) glasses [112].
Materials 11 00828 g009
Figure 10. Mossbauer spectra of nanocrystalline Sc 1.8 Eu 0.2 O 3 . The experimental data (dots) and the fit curve (full line) are shown along with the components associated to the C 3 i and C 2 sites (dashed line and full line, respectively) [117].
Figure 10. Mossbauer spectra of nanocrystalline Sc 1.8 Eu 0.2 O 3 . The experimental data (dots) and the fit curve (full line) are shown along with the components associated to the C 3 i and C 2 sites (dashed line and full line, respectively) [117].
Materials 11 00828 g010
Figure 11. The average isomer shift as a function of the measured lattice parameter, a, for a series of cubic Y 2 x Eu x O 3 samples. The solid line indicates the volume dependence [66].
Figure 11. The average isomer shift as a function of the measured lattice parameter, a, for a series of cubic Y 2 x Eu x O 3 samples. The solid line indicates the volume dependence [66].
Materials 11 00828 g011
Figure 12. Relationship between the isomer shift and the Eu–O bond length in some oxide crystals: 1—c-type Eu 2 O 3 (6); 2—EuBa 2 Cu 3 O 7 (8); 3—Eu 3 Ga 5 O 12 garnet (8); 4—EuAl 3 (BO 3 ) 4 huntite (9); 5—EuAlO 3 perovskite (12). The coordination number of Eu 3 + ion is shown in parentheses [114].
Figure 12. Relationship between the isomer shift and the Eu–O bond length in some oxide crystals: 1—c-type Eu 2 O 3 (6); 2—EuBa 2 Cu 3 O 7 (8); 3—Eu 3 Ga 5 O 12 garnet (8); 4—EuAl 3 (BO 3 ) 4 huntite (9); 5—EuAlO 3 perovskite (12). The coordination number of Eu 3 + ion is shown in parentheses [114].
Materials 11 00828 g012
Table 1. Data for Mössbauer isomer shifts of Eu in binary compounds. Data from [11].
Table 1. Data for Mössbauer isomer shifts of Eu in binary compounds. Data from [11].
Compound with Divalent EuIsomer Shift in mm/sCompound with Trivalent EuIsomer Shift in mm/s
EuF 2 13.58 EuF 3 0 (standard)
EuBr 2 13.43 EuCl 3 0.05
EuCl 2 13.10 EuBr 3 0.17
EuI 2 12.50 EuI 3 0.29
EuO 12.28 Eu 2 O 3 1.01
Table 2. Data for isomer shifts of Eu 2 + at room temperature with respect to EuF 3 in different hosts. U—broad unresolved magnetic hyperfine structure, P—partially resolved magnetic hyperfine structure.
Table 2. Data for isomer shifts of Eu 2 + at room temperature with respect to EuF 3 in different hosts. U—broad unresolved magnetic hyperfine structure, P—partially resolved magnetic hyperfine structure.
HostIsomer Shift in mm/sHyperfine Structure for Diluted Eu 2 + References
ZBLAN and Cl glasses 13 U[91]
CaF 2 13.4 hyperfine field of 34 T[92]
ZrF 4 ·BaF 2 ·EuF 2 ·ThF 4 glass 14.18 ( g ) U[93]
ZBLAN glass 14.3 ( d ) U[94]
Sr 0.95 Eu 0.05 Al 12 O 19 9.7 [95]
Ba 0.95 Eu 0.05 Al 12 O 19 10.1 [95]
Ca 0.5 Eu 0.5 Al 12 O 19 11.2 [95]
Ca 0.95 Eu 0.05 Al 12 O 19 11.7 [95,96]
Eu 2 AlO 3.75 N 0.1 11.98 ( e ) [97]
BaMgAl 10 O 17 :Eu 12 ( f ) [3]
BaMgAl 10 O 17 :6%Eu 12.55 ( a , h , c ) [98]
EuAl 2 O 4 13.1 ( a ) [97]
EuMgAl 11 O 19 13.14 P[99]
Ba 0.5 Eu 0.5 MgAl 11 13.4 [100]
Lu 0.7 Eu 0.3 MgAl 11 O 19 13.47 P[99]
BaAl 2 Si 2 O 8 14 ( f ) U[45]
Sr 0.75 Eu 0.25 Al 12 O 19 14.2 ( h ) U[101]
B 2 O 3 glass 13 ( a ) hyperfine field of 35 T[102]
(0.7B 2 O 3 ·0.3Na 2 O) 0.85 ·(EuO) 0.15 glass 13.15 P[36]
(0.7B 2 O 3 ·0.3Na 2 O) 0.80 ·(EuO) 0.20 glass 13.15 P[36]
(0.7B 2 O 3 ·0.3Na 2 O) 0.70 ·(EuO) 0.30 glass 13.48 P[36]
(0.9B 2 O 3 ·0.1Na 2 O) 0.85 ·(EuO) 0.15 glass 13.55 P[36]
Eu 2 SiN 3 10.66 U[50]
CaS:Eu 11.5 [103]
EuS 11.7 [104]
Eu 2 VO 4 11.8 [105]
TiO 2 :Eu 12.04 ( b ) [106]
EuSi 2 O 2 N 2 12.3 ( f ) [52]
Eu 3 V 2 O 7 12.5 [105]
( a ) average over different sites; ( b ) converted from Eu 2 O 3 ; ( c ) 4.2 K; ( d ) 6 K; ( e ) 20 K; ( f ) 77/78 K; ( g ) increasing with decreasing temperature; ( h ) decreasing with increasing Eu concentration.
Table 3. Data for isomer shifts of Eu 3 + at room temperature with respect to EuF 3 in different hosts.
Table 3. Data for isomer shifts of Eu 3 + at room temperature with respect to EuF 3 in different hosts.
HostIsomer Shift in mm/sReferences
ZrF 4 ·BaF 2 ·EuF 2 ·ThF 4 glass0.06 ( g ) [93]
ZBLAN glass0 ( d ) [94]
ZBLAN and Cl glasses0[91]
(45 x )AlF 3 ·xAlPO 4 ·5EuF 3 ·30CaF 2 ·20BaF 2 0.05 [107]
Al 1.93 Eu 0.07 O 3 (mainly amorphous)1.11[108]
Eu 2 AlO 3.75 N 0.1 0.28 ( e ) [97]
Eu 0.1 Al 0.9 PO0.22[109]
La 0.7 Eu 0.3 MgAl 11 O 19 0.28 [99]
Sr 0.75 Eu 0.25 Al 12 O 19 0.3 ( h ) [101]
LaMgAl 11 O 19 0.35 [99]
BaMgAl 10 O 17 0.93 ( c ) [98]
SiO 2 ·Na 2 O·BaO·ZnO·Eu 2 O 3 1.07 ( b ) [110]
SiO 2 ·B 2 O 3 ·Na 2 O·Eu 2 O 3 1[111]
BaO·SiO 2 0.82[112]
50Na 2 O·44P 2 O 5 ·3Eu 2 O 3 ·3CeO 2 0.79 ( b ) [113]
CaO·SiO 2 0.76[112]
50SiO 2 ·25A1 2 0 3 ·25Eu 2 O 3 0.46[114]
(4ZnO) 0.975 ·3B 2 O 3 ·0.025Eu 2 O 3 0.45[115]
0.1Eu(PO 3 ) 3 ·0.9Zn(PO 3 ) 2 0.37[115]
0.1Eu(PO 3 ) 3 ·0.9Sr(PO 3 ) 2 0.36[115]
CaO·P 2 O 5 0.36[112]
BaO·P 2 O 5 0.32[112]
0.1Eu(PO 3 ) 3 ·0.9Pb(PO 3 ) 2 0.29[115]
(0.7B 2 O 3 ·0.3Na 2 O) 1 x ·(EuO) x 0[36]
In 1.9 Eu 0.1 O 3 1.38 ( a ) [66]
Lu 1.9 Eu 0.1 O 3 1.35 ( a ) [66]
Y 2 O 3 :Eu 3 + (nanoparticle)1.28 ( h ) [67,116]
Sc 1.8 Eu 0.2 O 3 1.28 ( a ) [117]
Sc 1.9 Eu 0.1 O 3 1.27 ( a ) [66]
Lu 1.8 Eu 0.2 O 3 1.25 ( a ) /1.23 ( a ) [117,118]
La 1.9 Eu 0.1 O 3 1.2[66]
Y 1.8 Eu 0.2 O 3 1.2[100]
Y 1.8 Eu 0.2 O 3 1.14 ( a ) [117]
Y 1.9 Eu 0.1 O 3 1.12 ( a ) [66]
Gd 1.9 Eu 0.1 O 3 (monoclinic)1.08 ( a ) [66]
Gd 1.9 Eu 0.1 O 3 (cubic)1.05 ( a ) [66]
Y 2 O 3 @Eu 3 + (core-shell)1.03[67,116]
Gd 1.8 Eu 0.2 O 3 (nanocrystalline)0.97[86,119]
Eu 3 V 2 O 7 0.6[105]
YVO 4 (nanoparticle)0.5 ( b ) [120]
Eu 2 VO 4 0.5 ( a ) [105]
Y 0.963 Eu 0.037 VO 4 0.49[100]
EuVO 4 (zircon) 0.06 [121]
EuVO 4 (scheelite) 0.55 [121]
YAG:Eu nanocrystal1.47 ( b ) [81]
BaCa 2 Y 6 O 12 :Eu1.0 ( a ) [53]
Eu 2 SiN 3 0.85[50]
Eu 2 Ti 2 O 7 0.8 ( b ) [87]
KY 0.735 Eu 0.265 W 2 O 8 0.68 ( f ) [122]
Y 1.95 Eu 0.05 WO 6 0.65 ( a , i ) [123]
CuLa 0.99 Eu 0.01 O 2 0.61 ( i ) [124]
6 TiO 2 :Eu0.6 ( h ) [125]
(BaGeO 3 ) 1 x y ·(Al 2 O 3 ) x ·(0.45CaF 2 ·0.55MgF 2 ) y 0.48[126]
TiO 2 :Eu0.38 ( b ) [106]
ZrO 2 :Eu0.3[127]
CaS:Eu 0.8 [103]
( a ) average over different sites; ( b ) converted from Eu 2 O 3 ; ( c ) 4.2 K; ( d ) 6 K; ( e ) 20 K; ( f ) 77/78 K; ( g ) increasing with decreasing temperature; ( h ) decreasing with increasing Eu concentration; ( i ) increasing with increasing Eu concentration.

Share and Cite

MDPI and ACS Style

Steudel, F.; Johnson, J.A.; Johnson, C.E.; Schweizer, S. Characterization of Luminescent Materials with 151Eu Mössbauer Spectroscopy. Materials 2018, 11, 828. https://doi.org/10.3390/ma11050828

AMA Style

Steudel F, Johnson JA, Johnson CE, Schweizer S. Characterization of Luminescent Materials with 151Eu Mössbauer Spectroscopy. Materials. 2018; 11(5):828. https://doi.org/10.3390/ma11050828

Chicago/Turabian Style

Steudel, Franziska, Jacqueline A. Johnson, Charles E. Johnson, and Stefan Schweizer. 2018. "Characterization of Luminescent Materials with 151Eu Mössbauer Spectroscopy" Materials 11, no. 5: 828. https://doi.org/10.3390/ma11050828

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