Next Article in Journal
2.0 μm Ultra Broadband Emission from Tm3+/Ho3+ Co-Doped Gallium Tellurite Glasses for Broadband Light Sources and Tunable Fiber Lasers
Next Article in Special Issue
Heat-Induced Transformation of Luminescent, Size Tuneable, Anisotropic Eu:Lu(OH)2Cl Microparticles to Micro-Structurally Controlled Eu:Lu2O3 Microplatelets
Previous Article in Journal
Preparation, Sintering Behavior and Consolidation Mechanism of Vanadium-Titanium Magnetite Pellets
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Luminescence Intensity Ratio Thermometry with Er3+: Performance Overview

by
Aleksandar Ćirić
*,
Tamara Gavrilović
and
Miroslav D. Dramićanin
*
Vinča Institute of Nuclear Sciences—National Institute of the Republic of Serbia, University of Belgrade, P.O. Box 522, 11001 Belgrade, Serbia
*
Authors to whom correspondence should be addressed.
Crystals 2021, 11(2), 189; https://doi.org/10.3390/cryst11020189
Submission received: 3 February 2021 / Revised: 9 February 2021 / Accepted: 10 February 2021 / Published: 14 February 2021
(This article belongs to the Special Issue Optical and Spectroscopic Properties of Rare-Earth-Doped Crystals)

Abstract

:
The figures of merit of luminescence intensity ratio (LIR) thermometry for Er3+ in 40 different crystals and glasses have been calculated and compared. For calculations, the relevant data has been collected from the literature while the missing data were derived from available absorption and emission spectra. The calculated parameters include Judd–Ofelt parameters, refractive indexes, Slater integrals, spin–orbit coupling parameters, reduced matrix elements (RMEs), energy differences between emitting levels used for LIR, absolute, and relative sensitivities. We found a slight variation of RMEs between hosts because of variations in values of Slater integrals and spin–orbit coupling parameters, and we calculated their average values over 40 hosts. The calculations showed that crystals perform better than glasses in Er3+-based thermometry, and we identified hosts that have large values of both absolute and relative sensitivity.

1. Introduction

The measurements of temperature, one of seven fundamental physical quantities, can be classified according to the nature of contact between the measurement object and instrument to invasive (where there is direct contact, e.g., thermocouples, thermistors), semi-invasive (where measuring object is altered in a way to enable contactless measurements), and non-invasive (where the temperature is estimated remotely, e.g., optical pyrometers) [1]. The first type necessarily perturbs the temperature of measurement objects which limits its use in microscopic objects. In addition, such approaches are difficult to implement on moving objects or in harsh environments, for example, in high-intensity electromagnetic fields, radioactive, or chemically challenging surroundings. Thus, the current market of thermometers, accounting for more than 80% of all sensors [2], demands methods that allow for remote or microscopic measurements. Among many perspective optical semi-invasive techniques, luminescence thermometry which uses thermographic phosphors has drawn the largest attention [3,4]. The thermographic phosphor probe can be incorporated within the measured object or on its surface, on macroscale to nanoscale sizes, or can be mounted on the surface of the fiber-optic cables and bring to proximity of measuring objects. Luminescent thermometry has found a range of valuable applications, from engineering to biomedical [5], and, currently, it is a widely researched topic with an exponentially increasing number of published research papers [6].
Presently, many types of materials are used for the construction of thermometry probes. These include rare-earth and transition metal activated phosphors, semiconductor quantum dots, organic dyes, and metal-organic complexes, carbon dots, and luminescent polymers. Among the rare-earth crystals are by far the most exploited type [5], usually exploited in the so-called luminescence intensity ratio (LIR, sometimes called fluorescence intensity ratio (FIR) or labeled as Δ) temperature read-out scheme that is based on the temperature-dependent intensity ratio of emissions from the thermally coupled excited levels of rare-earth ions.
The Er3+ is considered to be workhorse in LIR-based luminescence thermometry. LIR read-out is obtained with both downshifting and up-conversion emissions, the latter most frequently sensitized by Yb3+. The popularity of Er3+ in luminescence thermometry is a consequence of its efficient emission and ~700 cm−1 energy difference between its thermally coupled 4S3/2 and 2H11/2 levels. Such energy difference is ideal for thermometry in the physiological range of temperatures (30–50 °C) since it provides the maximal sensitivity of measurement in this range (~ 1.1%K−1 at 303 K). To increase sensitivity, in recent times, the emission from the Er3+ 4F7/2 level is used for LIR considering the higher energy difference between 4F7/2 and 4S3/2 levels compared to the energy difference between 2H11/2 and 4S3/2 levels [7]. This Er3+ LIR variant has perspective applications at high temperatures and greatly widens the sensor’s operating temperature range.
Since Er3+ can activate many hosts there are many investigations of luminescence thermometry using emissions of this ion. These studies are lengthy and cumbersome involving material synthesis and characterizations, measurements of emission spectra at various temperatures, complex data fitting and analysis, and evaluation of thermometric performance. Therefore, to alleviate this problem, the theoretical prediction of Er3+ thermometric performance in different hosts may be useful as a guide for the host selection.
Our previous research has demonstrated that LIR and figures of merit of luminescence thermometry can be predicted by a theoretical model that involves the famous Judd–Ofelt (JO) theory with the high matching to experimental data [6,8]. JO theory explains and predicts the intensities of the trivalent rare-earth ions’ (RE3+) f-f electronic transitions, and its parameters include all phenomenological mechanisms responsible for the line strengths observed in both absorption and emission spectra.
Here, we aimed to perform the theoretical analysis of Er3+ emissions involved in the LIR thermometry, linking the temperature sensing performance of materials (LIR absolute and relative sensitivities) with their composition and structural properties. For this extensive analysis we have selected 40 different materials in crystal and glass form, and collected the relevant data from the literature. The refractive index values were calculated from Sellmeier’s equation (where available). The energy level positions are given for each host. For 16 hosts, the reduced matrix elements (RMEs) are recalculated from the Slater integrals and spin–orbit (s–o) parameters. The relation between material properties, JO parameters, Slater integrals, and s–o parameters is provided.

2. Methods

2.1. Luminescence Intensity Ratio Method

Out of all the luminescence temperature read-out methods the LIR has been the most frequently investigated since it is simple, requires inexpensive equipment, and, most importantly, is ratiometric and self-referenced [7]. The temperature read-out is not affected by fluctuations in excitation power. The most interesting materials for LIR thermometry are trivalent lanthanides, as they have sharp emission lines and plenty to choose from, from ultraviolet (UV) to near-infrared (NIR) spectral regions. LIR is defined as the ratio of two emissions that varies with temperature [9]. In the case of two energy levels separated by < 2000 cm−1, it is said that they are thermally coupled, as the higher energy level can be effectively populated by the thermal energy. The ratio of populations of optical centers in the energetically higher (H) and lower (L) levels is then proportional to the Boltzmann distribution: NH/NL~exp(−ΔE/kT), where ΔE is the energy difference between the levels of interest, k = 0.695 cm−1 K−1 is the Boltzmann constant, and T is the temperature. As the intensities are proportional to the population, the equation for LIR for the Boltzmann thermometer is:
L I R ( T ) = I H ( T ) I L ( T ) = B × exp ( Δ E k T )
where B is the temperature invariant parameter that depends only on the host.
The thermometer’s performance is estimated by the absolute ( S a ) and relative ( S r ) sensitivities, and temperature resolution ( Δ T ) given by [10]:
S a = | L I R T | ,   S r = S a L I R · 100 % ,   Δ T = σ a S a = σ r S r
where σa and σr are the absolute and relative uncertainties in measurement of LIR, presented as standard deviations.
For the temperature dependence of LIR given by Equation (1), sensitivities have the following form:
S a = Δ E k T 2 B exp ( Δ E k T ) ,   S r = Δ E k T 2 · 100 %
The absolute sensitivity reaches maximum at T = Δ E / 2 k with the value of [8]:
S a m a x = 4 B k e 2 Δ E
where e = 2.718 is a number (the natural logarithm base).
The ideal situation for LIR is the Boltzmann luminescence thermometer since it is easily calibrated with the well-known and simple theory. According to Equation (3), the relative sensitivity only depends on the value of energy difference between thermalized energy levels. The choice of levels with the energy gap larger than 2000 cm−1 may result in the thermalization loss at low temperatures, and even around room temperatures, while the small energy gap gives small relative sensitivities. One should consider that for achieving the Boltzmann’s thermal equilibrium some other conditions must be fulfilled besides the suitable energy difference between the levels, as recently demonstrated by Geitenbeek et al. [11] and Suta et al. [12]. Furthermore, considering the adjacent energy levels of trivalent rare-earth used for thermometry, the largest energy gap is in the Eu3+ (between 5D1 and 5D0 levels) and is approximately 1750 cm−1.
Thus, the current research of LIR of a single emission center is aimed at increasing the relative sensitivity without the loss of thermalization (and deviation from the Boltzmann distribution). One recently demonstrated solution is the inclusion of the third, non-adjacent level, with higher energy, which is thermalized with the second level. If the first and second levels are thermalized, and second and third levels are thermalized, then the ratios of emission intensities of the first and third levels will follow the Boltzmann distribution, even if their separation is greater than stated above, see Figure 1 for the case of Er3+. The conventional LIR of Er3+ is equal to the ratio of emissions from 2H11/2 (~523 nm) and 4S3/2 (~542 nm) levels, which are separated by ~700 cm−1, thus giving the relative sensitivity of ~1.1% K−1. By observing intensities to the ground level from 4F7/2 (~485 nm) with the 4S3/2, it is evident that their relative change is much larger than with 2H11/2. This larger energy difference, according to Equation (3), ultimately results in more than a three-fold increase in relative sensitivity.

2.2. Judd–Ofelt Theory and Its Relevance for Luminescence Thermometry

The electronic configuration of trivalent erbium is that of Xenon plus the 11 electrons in the 4f shell, i.e., Er3+ has [Xe]4f11 electronic configuration. With only 3 electrons missing from the completely filled 4f shell, Er3+ shares the same LS terms and LSJ levels as the Nd3+ who has 3 electrons in the 4f shell. The transitions from one to another level are followed by the reception or release of energy. The probability for such phenomena for a given set of initial and final levels is given by the wavefunctions and the appropriate moment operator. The exchange of energy in the intra-configurational 4f transitions with the highest intensity is of induced electric dipole and magnetic dipole types [13]. What was puzzling only half a century ago was the origin of these “electric dipole” interactions, as they were clearly and strictly forbidden by the Parity selection rule, also known as the Laporte rule. The solution to this problem came in 1962 in the papers simultaneously published by Judd [14] and Ofelt [15], to what is latter known as the Judd–Ofelt theory (JO). For the sake of brevity, it will not be explained here, and the reader is instead referred to the excellent references [16,17,18]; however, we will touch on several basics that are the most relevant for the present research.
In RE3+ (trivalent rare-earth) ions in general, the electrostatic (He) and s–o (Hso) interactions between 4f electrons are dominant and are of the approximately same magnitude, thus the Hamiltonian can be given approximately as H = He + Hso [17].
The electrostatic Hamiltonian can be reduced to the electron-electron repulsion form [17], which can be split into its radial (Fk) and angular (fk) parts:
H e = 1 4 π ε 0 i < j 11 e 2 r i j = k F k f k
The radial parameters are the Slater integrals given by [19]:
F k ( 4 f ) = e 2 4 π ε 0 0 r < k r > k + 1 R 4 f 2 ( r i ) R 4 f 2 ( r j ) d r i d r j
where r> is greater and r< is smaller than ri and rj, and R are the radial parts of the wavefunction. The Slater integrals can be evaluated by the Hartree-Fock method; however, it does not provide accurate results, and it is best to obtain them semi-empirically by adjusting them to the experimentally observed energies of 4f levels [20].
Hso mixes all states that have the same J quantum number, and it is proportional to the s–o coupling parameter, ζ, which is further proportional to the number of electrons within the 4f shell. Er3+ has a relatively high value of ζ in comparison with other trivalent lanthanide ions, providing a large mixing of states [21]. In this intermediate coupling approximation scheme, the wavefunctions are expressed as a linear combination of all other states in the configuration with the same J quantum number [17,18,22,23]:
| 4 f 11 S L J = i c i | 4 f 11 S L J ,   i c i 2 = 1
As the 4f electrons are shielded by the outer higher-energy electrons the crystal field (CF) introduces only a perturbation to the Hamiltonian [16]. Nevertheless, that perturbation weakens the already mentioned Laporte (parity) selection rule that forbids the ED transitions within the configuration. The 4f–4f transitions of electric dipole (ED) type become allowed and are known as the induced ED [24]. The radiative transition probability for such spontaneous emission is then equal to [25]:
A S L J S L J = 64 π 4 ν ˜ S L J S L J 3 3 h ( 2 J + 1 ) ( χ E D D E D + χ M D D M D )
or for the purely induced ED emission (MD is an abbreviation for the magnetic dipole):
A S L J S L J = 64 π 4 ν ˜ S L J S L J 3 3 h ( 2 J + 1 ) χ E D D E D
where h = 6.626 × 10−27 erg∙s is the Planck’s constant. Χ is the local field correction, ν ˜ S L J S L J is the emission barycenter energy, and D is the dipole strength given in esu2 cm2 units. The emission barycenter is [26]:
ν ˜ S L J S L J =   ν S L J S L J i S L J S L J ( ν ) d ν   i S L J S L J ( ν ) d ν
where i is the intensity at a given energy. The local field correction for ED emission is given by [27]:
χ E D = n ( n 2 + 2 ) 2 9
where n is the refractive index that should be given at the wavelength of the barycenter of the emission. It can be calculated from the Sellmeier’s equation for a given material, which is given in the form [28,29]:
n ( λ ) = 1 + i = 1 3 B i λ 2 λ 2 C i
In the JO scheme, the ED strength is given by [26]:
D E D S L J S L J = e 2 λ = 2 , 4 , 6 Ω λ U S L J S L J λ
where e = 4.803 × 1010 esu, U S L J S L J λ is the abbreviation for squared RMEs | 4 f 11 S L J U λ 4 f 11 S L J | 2 , which in turn can be calculated from the Slater integrals and the s–o coupling parameter. Ωλ are the JO intensity parameters, obtained semi-empirically or by the ab initio calculations (from the crystal-field parameters).
The integrated emission intensity for the transition SLJSLJ′ is given by [30,31]:
I S L J S L J =   i S L J S L J ( ν ˜ ) d ν ˜ = h ν ˜ S L J S L J N S L J A S L J S L J
or without the if the spectrum is recorded in counts instead of power units [32].
LIR of two emissions from the thermally coupled levels is then given by:
L I R ( T ) = I H I L = ν H N H A H ν L N L A L
where IH/L are the integrated intensities from the higher and lower level, respectively (without νH/νL if recorded in counts).
According to the Boltzmann distribution, the optical center population is given by:
N H N L = g H g L exp ( Δ E k T )
where g = 2J + 1 are the degeneracies of the selected levels.
Equation (15) can be rewritten as Equation (1) where B is the temperature invariant parameter that is given by:
B = ν H g H A H ν L g L A L
or if the intensities are recorded in counts instead of power units, without the νH/νL. As we have demonstrated in our previous article [8], by inserting Equation (8) into Equation (17), the LIR, the absolute sensitivity (and everything related to it) and temperature resolution can be predicted by JO parameters, as the B parameter can be obtained from:
B = ( ν ˜ H ν ˜ L ) 4 χ E D H D E D H + χ M D H D M D H χ E D L D E D L + χ M D L D M D L
or in the case of the pure ED transitions:
B = ( ν ˜ H ν ˜ L ) 4 χ E D H D E D H χ E D L D E D L
For the case of spectra recorded in counts, νH/νL should be to the power of 3.
The shielding of 4f electrons by electrons from outer orbitals ensures that the RE3+ spectra are featured by sharp peaks whose energies are almost host-independent. This is reflected in the almost host invariant reduced matrix elements. However, as the Slater parameters deviate significantly in Er3+, using such approximation may introduce significant errors. For the analysis of this type, it is more accurate to use the reduced matrix elements that are calculated from Slater integrals and s–o coupling parameters, which are calculated semi-empirically from the positions of the energy levels. Analogously, the small variations in energy level positions may provide significant variations in energy level differences, and thus large deviations in absolute and relative sensitivities. Finally, the small differences in refractive index become enormous when they propagate in the local field correction coefficient and, thus, it is of utmost importance to use accurate values. In this study, the observed levels are energetically very close, thus, it is a good approximation to consider the refractive index as wavelength-independent; however, the exact method is always preferred.

3. Results and Discussion

3.1. Calculations of Er3+ Radiative Properties in Different Hosts

For the study, we have selected 40 different hosts doped with Er3+ (Table 1), from the literature that contained the most complete set of data needed for the analysis presented in this paper. As the JO parametrization is traditionally performed semi-empirically from the absorption spectrum, powders and non-transparent materials are not included in this analysis.
In the 3rd column, Table 2 gives the energies of the 4S3/2, 2H11/2, and 4F7/2 levels used for the two LIRs that will be theoretically investigated. As stated in the introduction, this is important in the estimation of the thermometric figures of merit, and it is linked to the Slater integrals and s–o parameters. The table also includes the Slater integrals and s–o coupling parameters of the 16 out of the 40 hosts, and the JO intensity parameters for all the hosts, taken from references Table 1.
Figure 2 presents the variation of Slater integrals and s–o coupling parameters in those 16 hosts. Although there are no large differences in parameters between the crystals and glasses, there are certain trends that may be observed by the type of compound. Deviations in parameters’ values from host to host can be large, so the use of Carnall or Weber tables [22,35] for Er3+ RMEs can introduce large errors in the later calculations.
Figure 3 presents the JO parameters as given in Table 2. Glass hosts have smaller values of JO parameters than crystals, on average. When crystals are analyzed, the largest values of Ω2 parameter are found in tungstates and molybdates, while the smallest values are in garnets, phosphates, silicates, and oxysulfides. Ω6 are expectedly higher in fluorides, phosphates, and silicates. In glasses, borate glasses have lower Ω2, while phosphate glasses have higher Ω2. Ω6 is on average higher in phosphate glasses. No clear correlation could be given for the Ω4 parameter in crystals or glasses.
The squared RMEs for each transition investigated for LIR are given in Table 3. This list can be used beyond the scope of this paper for accurate calculations of JO parameters. The deviations from the average RME values are given in Figure 4, and they are large for the 2H11/24I15/2 transition. Thus, the use of Carnall’s or Weber’s values [22,35] might introduce significant errors in the JO parameters estimation, as the RMEs were calculated for the LaF3 and YAlO3, respectively. The average RMEs values calculated from Table 3 are given in Table 4, together with the deviations from the values by Carnall and Weber. The refractive index values taken from the corresponding references are also listed in Table 3. If Sellmeier’s equation is given, the refractive index is calculated at the wavelength of the emission. From the refractive index value, the local field correction is calculated according to Equation (11). The induced ED strengths (the last column of Table 3) are calculated for each transition and for each using JO parameters from Table 2 and local field corrections and RMEs from Table 3.

3.2. Calculations of LIR Parameters

For this theoretical analysis, two Er3+-based LIRs are considered, the traditional LIR that uses the temperature-dependent ratio of emissions from 4S3/2 and 2H11 levels, and the relatively novel concept that uses the temperature-dependent ratio emissions from 4S3/2 and 4F7/2 levels. Table 5 provides the energy differences between 4S3/2 and 2H11 and 4S3/2 and 4F7/2 that are used to calculate the room-temperature-relative sensitivities for each host using Equation (2). The temperature invariant B parameters are calculated from the data in Table 3 using Equation (19) (version for spectra recorded in counts). Then, using Equations (2)–(4) and calculated B values, it was possible to derive the LIR’s absolute sensitivity, the maximal absolute sensitivity value, and the temperature at which maximal absolute sensitivity occurs.
The relation between relative and absolute sensitivities of traditional LIR (that uses Er3+ emissions from 2H11/2 and 4S3/2 levels) for different hosts is presented in Figure 5a–c. As a rule of thumb, the higher the sensitivity value the better is the performance of thermometry. From Figure 5a, one can see that glasses tend to perform slightly weaker than crystals, on average. Figure 5b compares the LIR performance of different crystals. Fluorides’, garnets’, phosphates’, and silicates’ performances are worse than for other hosts. The best results are obtained with simple oxides, vanadates, niobates, molybdates, and tungstates. Figure 5c illustrates the performances of only glass hosts. Even the number of hosts in this set is rather small, it is possible to observe that Er3+ activated borate glasses perform worse than other glasses. Fluorophosphate glasses show high relative sensitivities, but somewhat small absolute sensitivities. The best combination of sensitivities is achieved in PbO-PbF2 glass. Similar conclusions can be drawn for the novel LIR type (that uses Er3+ emissions from 4F7/2 and 4S3/2 levels), Figure 5d–f. Among different glasses, tellurite-fluoride glasses show the best performance. For crystals, the situation is almost equivalent to that of traditional LIR.
Figure 6a–c gives the relation between relative sensitivity and absolute sensitivity at the temperature at which the absolute sensitivity has its maximum for the traditional LIR, while Figure 6d–f show the same relationship for the novel type LIR. Analogous conclusions can be drawn as in the previous analysis (Figure 5). Among glass hosts, tellurite-fluoride, tungstate, and molybdate glasses show the best performances. Among crystals, the performance trend is almost the same, but the NaY(MoO4)2 shows the worst performance at elevated temperatures. The best overall performer is LiLa(WO4)2.
As a limit of the study, we must note that the values of the energy levels, Slater integrals and s–o parameters, refractive index values, and JO parameters are taken from literature, so one cannot estimate the level of their accuracy. The extreme outliers are to be taken with caution.

4. Conclusions

The conventional thermometric characterizations are lengthy, complicated, and expensive. Given that there is an infinite number of possible hosts and doping concentrations of luminescent activators, the guidelines in selecting the appropriate material are important, and they can be provided by the Judd–Ofelt thermometric model which predicts thermometric figures of merit from its 3 intensity parameters.
Er3+ deserves special attention in luminescence thermometry. It features LIR between 2H11/2 and 4S3/2 levels with energy separation of ~700 cm−1, and a recently introduced LIR between 4F7/2 and 4S3/2 levels, whose higher energy separation allows for up to 3× larger relative sensitivity. The performances of 40 various crystals and glasses were predicted by the Judd–Ofelt thermometric model, and guidelines were set to aid the search for the best phosphor for LIR thermometry.
It was demonstrated that the Slater integrals and s–o coupling parameters significantly vary from host to host so that their values should not be adopted from other hosts. Consequently, for Er3+, the squared reduced matrix elements also significantly vary between hosts (especially for the 2H11/24I15/2 transition). Therefore, RMEs from frequently used Carnall or Weber tables should be replaced by the average RMEs for the three transitions that are used in these LIR read-out schemes, if the exact RMEs cannot be obtained. This will allow for the improved precision in the prediction of thermometric sensor performances, as well as for the improved Judd–Ofelt parametrization of Er3+ doped compounds.

Author Contributions

Conceptualization, A.Ć. and M.D.D.; methodology, A.Ć. and M.D.D.; validation, A.Ć.; formal analysis, A.Ć.; investigation, A.Ć. and T.G.; resources, A.Ć. and T.G.; data curation, A.Ć.; writing—original draft preparation, A.Ć. and M.D.D.; writing—review and editing, A.Ć. and M.D.D.; visualization, A.Ć.; supervision, M.D.D.; project administration, M.D.D.; funding acquisition, M.D.D. All authors have read and agreed to the published version of the manuscript.

Funding

This research was supported by the NATO Science for Peace and Security Programme under award id. [G5751] and by the Ministry of Education, Science, and Technological Development of the Republic of Serbia.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data are available from Aleksandar Ćirić upon reasonable request.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Childs, P.R.N.; Greenwood, J.R.; Long, C.A. Review of temperature measurement. Rev. Sci. Instrum. 2000, 71, 2959–2978. [Google Scholar] [CrossRef] [Green Version]
  2. Gschwend, P.M.; Starsich, F.H.L.; Keitel, R.C.; Pratsinis, S.E. Nd3+-Doped BiVO4 luminescent nanothermometers of high sensitivity. Chem. Commun. 2019, 55, 7147–7150. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  3. Wang, X.; Wolfbeis, O.S.; Meier, R.J. Luminescent probes and sensors for temperature. Chem. Soc. Rev. 2013, 42, 7834. [Google Scholar] [CrossRef] [PubMed]
  4. Dramićanin, M.D. Sensing temperature via downshifting emissions of lanthanide-doped metal oxides and salts. A review. Methods Appl. Fluoresc. 2016, 4, 42001. [Google Scholar] [CrossRef] [Green Version]
  5. Dramićanin, M.D. Luminescence Thermometry, Methods, Materials and Applications; Woodhead Publishing: Sawston, UK, 2018; ISBN 9780081020296. [Google Scholar]
  6. Ćirić, A.; Zeković, I.; Medić, M.; Antić, Ž.; Dramićanin, M.D. Judd-Ofelt modelling of the dual-excited single band ratiometric luminescence thermometry. J. Lumin. 2020, 225. [Google Scholar] [CrossRef]
  7. Dramićanin, M.D. Trends in luminescence thermometry. J. Appl. Phys. 2020, 128, 40902. [Google Scholar] [CrossRef]
  8. Ćirić, A.; Stojadinović, S.; Dramićanin, M.D. An extension of the Judd-Ofelt theory to the field of lanthanide thermometry. J. Lumin. 2019, 216. [Google Scholar] [CrossRef]
  9. Wade, S.A.; Collins, S.F.; Baxter, G.W. Fluorescence intensity ratio technique for optical fiber point temperature sensing. J. Appl. Phys. 2003, 94, 4743. [Google Scholar] [CrossRef]
  10. Ćirić, A.; Stojadinović, S.; Dramićanin, M.D. Custom-built thermometry apparatus and luminescence intensity ratio thermometry of ZrO2:Eu3+ and Nb2O5:Eu3+. Meas. Sci. Technol. 2019, 30, 45001. [Google Scholar] [CrossRef]
  11. Geitenbeek, R.G.; De Wijn, H.W.; Meijerink, A. Non-Boltzmann Luminescence in NaYF4:Eu3+: Implications for Luminescence Thermometry. Phys. Rev. Appl. 2018, 10, 1. [Google Scholar] [CrossRef] [Green Version]
  12. Suta, M.; Antić, Ž.; Ðorđević, V.; Kuzman, S.; Dramićanin, M.D.; Meijerink, A. Making Nd3+ a Sensitive Luminescent Thermometer for Physiological Temperatures—An Account of Pitfalls in Boltzmann Thermometry Making Nd3+ a Sensitive Luminescent Thermometer for Physiological Temperatures—An Account of Pitfalls in Boltzmann Thermomet. Nanomaterials 2020, 10, 543. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  13. Smentek, L. Judd-Ofelt Theory—The Golden (and the Only One) Theoretical Tool of f-Electron Spectroscopy. Comput. Methods Lanthan. Actin. Chem. 2015, 241–268. [Google Scholar] [CrossRef]
  14. Judd, B.R. Optical Absorption Intensities of Rare-Earth Ions. Phys. Rev. 1962, 127, 750–761. [Google Scholar] [CrossRef]
  15. Ofelt, G.S. Intensities of Crystal Spectra of Rare-Earth Ions. J. Chem. Phys. 1962, 37, 511–520. [Google Scholar] [CrossRef]
  16. Görller-Walrand, C.; Binnemans, K. Chapter 167 Spectral intensities of f-f transitions; Elsevier: Amsterdam, The Netherlands, 1998; Volume 25, pp. 101–264. [Google Scholar]
  17. Hehlen, M.P.; Brik, M.G.; Krämer, K.W. 50th anniversary of the Judd–Ofelt theory: An experimentalist’s view of the formalism and its application. J. Lumin. 2013, 136, 221–239. [Google Scholar] [CrossRef]
  18. Walsh, B.M. Judd-Ofelt theory: Principles and practices brian m. walsh. Int. Sch. At. Mol. Spectrosc. 2006, 403–433. [Google Scholar] [CrossRef]
  19. Reisfeld, R. Optical Properties of Lanthanides in Condensed Phase, Theory and Applications. AIMS Mater. Sci. 2015, 2, 37–60. [Google Scholar] [CrossRef]
  20. Hamm, P.; Helbing, J.; Liu, G.; Jacquier, B.; Thermo, S.; Swart, H.C.; Terblans, J.J.; Ntwaeaborwa, O.M.; Kroon, R.E.; Coetsee, E.; et al. Spectroscopic Properties of Rare Earths in Optical Materials; Springer Series in Materials Science; Hull, R., Parisi, J., Osgood, R.M., Warlimont, H., Liu, G., Jacquier, B., Eds.; Springer: Berlin/Heidelberg, Germany, 2005; Volume 83, ISBN 3-540-23886-7. [Google Scholar]
  21. Edelstein, N.M. Electronic Structure of f-Block Compounds. In Organometallics of the f-Elements; Springer Netherlands: Dordrecht, The Netherlands, 1979; pp. 37–79. [Google Scholar]
  22. Carnall, W.T.; Crosswhite, H.; Crosswhite, H.M. Energy Level Structure and Transition Probabilities in the Spectra of the Trivalent Lanthanides in LaF3; Argonne: Lemont, IL, USA, 1978.
  23. Weber, M.J.; Varitimos, T.E.; Matsinger, B.H. Optical intensities of rare-earth ions in yttrium orthoaluminate. Phys. Rev. B 1973, 8, 47–53. [Google Scholar] [CrossRef]
  24. Tanner, P.A. Some misconceptions concerning the electronic spectra of tri-positive europium and cerium. Chem. Soc. Rev. 2013, 42, 5090. [Google Scholar] [CrossRef] [PubMed]
  25. Binnemans, K. Interpretation of europium(III) spectra. Coord. Chem. Rev. 2015, 295, 1–45. [Google Scholar] [CrossRef] [Green Version]
  26. Ćirić, A.; Stojadinović, S.; Brik, M.G.; Dramićanin, M.D. Judd-Ofelt parametrization from emission spectra: The case study of the Eu3+ 5D1 emitting level. Chem. Phys. 2020, 528. [Google Scholar] [CrossRef]
  27. Ćirić, A.; Stojadinović, S.; Sekulić, M.; Dramićanin, M.D. JOES: An application software for Judd-Ofelt analysis from Eu3+ emission spectra. J. Lumin. 2019, 205, 351–356. [Google Scholar] [CrossRef]
  28. Ćirić, A.; Stojadinović, S.; Dramićanin, M.D. Temperature and concentration dependent Judd-Ofelt analysis of Y2O3:Eu3+ and YVO4:Eu3+. Phys. B Condens. Matter 2020, 579. [Google Scholar] [CrossRef]
  29. Preda, E.; Stef, M.; Buse, G.; Pruna, A.; Nicoara, I. Concentration dependence of the Judd–Ofelt parameters of Er3+ ions in CaF2 crystals. Phys. Scr. 2009, 79, 035304. [Google Scholar] [CrossRef]
  30. Carlos, L.D.; De Mello Donegá, C.; Albuquerque, R.Q.; Alves, S.; Menezes, J.F.S.; Malta, O.L. Highly luminescent europium(III) complexes with naphtoiltrifluoroacetone and dimethyl sulphoxide. Mol. Phys. 2003, 101, 1037–1045. [Google Scholar] [CrossRef]
  31. dos Santos, B.F.; dos Santos Rezende, M.V.; Montes, P.J.R.; Araujo, R.M.; dos Santos, M.A.C.; Valerio, M.E.G. Spectroscopy study of SrAl2O4:Eu3+. J. Lumin. 2012, 132, 1015–1020. [Google Scholar] [CrossRef]
  32. Suta, M.; Meijerink, A. A Theoretical Framework for Ratiometric Single Ion Luminescent Thermometers—Thermodynamic and Kinetic Guidelines for Optimized Performance. Adv. Theory Simul. 2020, 3, 2000176. [Google Scholar] [CrossRef]
  33. Villanueva-Delgado, P.; Biner, D.; Krämer, K.W. Judd–Ofelt analysis of β-NaGdF4:Yb3+,Tm3+ and β-NaGdF4:Er3+ single crystals. J. Lumin. 2017, 189, 84–90. [Google Scholar] [CrossRef]
  34. Krupke, W.F.; Gruber, J.B. Energy Levels of Er3+ in LaF3 and Coherent Emission at 1.61 μm. J. Chem. Phys. 1964, 41, 1225–1232. [Google Scholar] [CrossRef]
  35. Weber, M.J. Probabilities for Radiative and Nonradiative Decay of Er3+ in LaF3. Phys. Rev. 1967, 157, 262–272. [Google Scholar] [CrossRef]
  36. Sardar, D.K.; Bradley, W.M.; Perez, J.J.; Gruber, J.B.; Zandi, B.; Hutchinson, J.A.; Trussell, C.W.; Kokta, M.R. Judd–Ofelt analysis of the Er3+ (4f11) absorption intensities in Er3+ −doped garnets. J. Appl. Phys. 2003, 93, 2602–2607. [Google Scholar] [CrossRef]
  37. Buddhudu, S.; Bryant, F. Optical transitions of Er3+:La2O2S and Er3+:Y2O2S. J. Less Common Met. 1989, 147, 213–225. [Google Scholar] [CrossRef]
  38. Morrison, C.A.; Leavitt, R.P. Chapter 46 Spectroscopic properties of triply ionized. In Handbook on the Physics and Chemistry of Rare Earths; Elsevier: Amsterdam, The Netherlands, 1982; Volume 5, pp. 461–692. [Google Scholar]
  39. Moustafa, S.Y.; Sahar, M.R.; Ghoshal, S.K. Spectroscopic attributes of Er3+ ions in antimony phosphate glass incorporated with Ag nanoparticles: Judd-Ofelt analysis. J. Alloys Compd. 2017, 712, 781–794. [Google Scholar] [CrossRef]
  40. Kaminskii, A.A.; Mironov, V.S.; Kornienko, A.; Bagaev, S.N.; Boulon, G.; Brenier, A.; Di Bartolo, B. New laser properties and spectroscopy of orthorhombic crystals YAlO3:Er3+. Intensity luminescence characteristics, stimulated emission, and full set of squared reduced-matrix elements |⧼α[SL]J| |U(t)||α′[S′ L′]J′⧽|2 for Er3+ Ions. Phys. Status Solidi 1995, 151, 231–255. [Google Scholar] [CrossRef]
  41. Chen, C.Y.; Sibley, W.A.; Yeh, D.C.; Hunt, C.A. The optical properties of Er3+ and Tm3+ in KCaF3 crystal. J. Lumin. 1989, 43, 185–194. [Google Scholar] [CrossRef]
  42. Weber, M.J. Radiative and Multiphonon Relaxation of Rare-Earth Ions in Y2O3. Phys. Rev. 1968, 171, 283–291. [Google Scholar] [CrossRef]
  43. Kisliuk, P.; Krupke, W.F.; Gruber, J.B. Spectrum of Er3+ in Single Crystals of Y2O3. J. Chem. Phys. 1964, 40, 3606–3610. [Google Scholar] [CrossRef]
  44. Capobianco, J.A.; Kabro, P.; Ermeneux, F.S.; Moncorgé, R.; Bettinelli, M.; Cavalli, E. Optical spectroscopy, fluorescence dynamics and crystal-field analysis of Er3+ in YVO4. Chem. Phys. 1997, 214, 329–340. [Google Scholar] [CrossRef]
  45. Reisfeld, R.; Katz, G.; Spector, N.; Jørgensen, C.K.; Jacoboni, C.; De Pape, R. Optical transition probabilities of Er3+ in fluoride glasses. J. Solid State Chem. 1982, 41, 253–261. [Google Scholar] [CrossRef]
  46. Souriau, J.C.; Borel, C.; Wyon, C.; Li, C.; Moncorgé, R. Spectroscopic properties and fluorescence dynamics of Er3+ and Yb3+ in CaYAlO4. J. Lumin. 1994, 59, 349–359. [Google Scholar] [CrossRef]
  47. Jamalaiah, B.C.; Suhasini, T.; Rama Moorthy, L.; Janardhan Reddy, K.; Kim, I.-G.; Yoo, D.-S.; Jang, K. Visible and near infrared luminescence properties of Er3+-doped LBTAF glasses for optical amplifiers. Opt. Mater. 2012, 34, 861–867. [Google Scholar] [CrossRef]
  48. Renuka Devi, A.; Jayasankar, C.K. Optical properties of Er3+ ions in lithium borate glasses and comparative energy level analyses of Er3+ ions in various glasses. J. Non. Cryst. Solids 1996, 197, 111–128. [Google Scholar] [CrossRef]
  49. Merino, R.I.; Orera, V.M.; Cases, R.; Chamarro, M.A. Spectroscopic characterization of Er3+ in stabilized zirconia single crystals. J. Phys. Condens. Matter 1991, 3, 8491–8502. [Google Scholar] [CrossRef]
  50. Amin, J.; Dussardier, B.; Schweizer, T.; Hempstead, M. Spectroscopic analysis of Er3+ transitions in lithium niobate. J. Lumin. 1996, 69, 17–26. [Google Scholar] [CrossRef]
  51. Nachimuthu, P.; Jagannathan, R. Judd-Ofelt Parameters, Hypersensitivity, and Emission Characteristics of Ln3+ (Nd3+, Ho3+, and Er3+) Ions Doped in PbO-PbF2 Glasses. J. Am. Ceram. Soc. 2004, 82, 387–392. [Google Scholar] [CrossRef]
  52. Shinn, M.D.; Sibley, W.A.; Drexhage, M.G.; Brown, R.N. Optical transitions of Er3+ ions in fluorozirconate glass. Phys. Rev. B 1983, 27, 6635–6648. [Google Scholar] [CrossRef]
  53. Swapna, K.; Mahamuda, S.; Venkateswarlu, M.; Srinivasa Rao, A.; Jayasimhadri, M.; Shakya, S.; Prakash, G.V. Visible, Up-conversion and NIR (~1.5μm) luminescence studies of Er3+ doped Zinc Alumino Bismuth Borate glasses. J. Lumin. 2015, 163, 55–63. [Google Scholar] [CrossRef]
  54. Moorthy, L.R.; Jayasimhadri, M.; Saleem, S.A.; Murthy, D.V.R. Optical properties of Er3+-doped alkali fluorophosphate glasses. J. Non. Cryst. Solids 2007, 353, 1392–1396. [Google Scholar] [CrossRef]
  55. Sardar, D.K.; Gruber, J.B.; Zandi, B.; Hutchinson, J.A.; Trussell, C.W. Judd–Ofelt analysis of the Er3+(4f11) absorption intensities in phosphate glass: Er3+, Yb3+. J. Appl. Phys. 2003, 93, 2041–2046. [Google Scholar] [CrossRef]
  56. Lalla, E.A.; Konstantinidis, M.; De Souza, I.; Daly, M.G.; Martín, I.R.; Lavín, V.; Rodríguez-Mendoza, U.R. Judd-Ofelt parameters of RE3+-doped fluorotellurite glass (RE3+ = Pr3+, Nd3+, Sm3+, Tb3+, Dy3+, Ho3+, Er3+, and Tm3+). J. Alloys Compd. 2020, 845, 156028. [Google Scholar] [CrossRef]
  57. Piao, R.; Wang, Y.; Zhang, Z.; Zhang, C.; Yang, X.; Zhang, D. Optical and Judd-Ofelt spectroscopic study of Er3+-doped strontium gadolinium gallium garnet single-crystal. J. Am. Ceram. Soc. 2018, jace.16114. [Google Scholar] [CrossRef]
  58. Che, Y.; Zheng, F.; Dou, C.; Yin, Y.; Wang, Z.; Zhong, D.; Sun, S.; Teng, B. A promising laser crystal Er3+:YPO4 with intense multi-wavelength emission characteristics. J. Alloys Compd. 2020, 157854. [Google Scholar] [CrossRef]
  59. Huy, B.T.; Sengthong, B.; Van Do, P.; Chung, J.W.; Ajith Kumar, G.; Quang, V.X.; Dao, V.-D.; Lee, Y.-I. A bright yellow light from a Yb3+,Er3+ -co-doped Y2SiO5 upconversion luminescence material. RSC Adv. 2016, 6, 92454–92462. [Google Scholar] [CrossRef]
  60. Pan, Y.; Gong, X.H.; Chen, Y.J.; Lin, Y.F.; Huang, J.H.; Luo, Z.D.; Huang, Y.D. Polarized spectroscopic properties of Er3+:BaGd2(MoO4)4 crystal. Opt. Mater. 2012, 34, 1143–1147. [Google Scholar] [CrossRef]
  61. Lu, X.; You, Z.; Li, J.; Zhu, Z.; Jia, G.; Wu, B.; Tu, C. The optical properties of Er3+ doped NaY(MoO4)2 crystal for laser applications around 1.5 μm. J. Alloys Compd. 2006, 426, 352–356. [Google Scholar] [CrossRef]
  62. Lu, H.; Gao, Y.; Hao, H.; Shi, G.; Li, D.; Song, Y.; Wang, Y.; Zhang, X. Judd-Ofelt analysis and temperature dependent upconversion luminescence of Er3+/Yb3+ codoped Gd2(MoO4)3 phosphor. J. Lumin. 2017, 186, 34–39. [Google Scholar] [CrossRef]
  63. Huang, X.; Wang, G. Growth and optical characteristics of Er3+:LiLa(MoO4)2 crystal. J. Alloys Compd. 2009, 475, 693–697. [Google Scholar] [CrossRef]
  64. Huang, X.Y.; Lin, Z.B.; Zhang, L.Z.; Wang, G.F. Spectroscopic characteristics of Er3+/Yb3+:LiLa(WO4)2 crystal. Mater. Res. Innov. 2008, 12, 94–97. [Google Scholar] [CrossRef]
  65. Kuleshov, N.V.; Lagatsky, A.A.; Podlipensky, A.V.; Mikhailov, V.P.; Kornienko, A.A.; Dunina, E.B.; Hartung, S.; Huber, G. Fluorescence dynamics, excited-state absorption, and stimulated emission of Er3+ in KY(WO4)2. J. Opt. Soc. Am. B 1998, 15, 1205. [Google Scholar] [CrossRef]
  66. Carnall, W.T.; Fields, P.R.; Rajnak, K. Electronic Energy Levels in the Trivalent Lanthanide Aquo Ions. I. Pr3+, Nd3+, Pm3+, Sm3+, Dy3+, Ho3+, Er3+, and Tm3+. J. Chem. Phys. 1968, 49, 4424–4442. [Google Scholar] [CrossRef]
  67. Amotchkina, T.; Trubetskov, M.; Hahner, D.; Pervak, V. Characterization of e-beam evaporated Ge, YbF3, ZnS, and LaF3 thin films for laser-oriented coatings. Appl. Opt. 2020, 59, A40. [Google Scholar] [CrossRef]
  68. Sell, J.A.; Fong, F.K. Oscillator strength determination in LaCl3:Pr3+ by photon upconversion. J. Chem. Phys. 1975, 62, 4161–4164. [Google Scholar] [CrossRef]
  69. Imanaga, S.; Yokono, S.; Hoshina, T. Cooperative absorption in Eu2O2S. J. Lumin. 1978, 16, 77–87. [Google Scholar] [CrossRef]
  70. Nigara, Y. Measurement of the Optical Constants of Yttrium Oxide. Jpn. J. Appl. Phys. 1968, 7, 404–408. [Google Scholar] [CrossRef]
  71. Shi, H.-S.; Zhang, G.; Shen, H.-Y. Measurement of principal refractive indices and the thermal refractive index coefficients of yttrium vanadate. J. Synth. Cryst. 2001, 30, 85–88. [Google Scholar]
  72. Pirzio, F.; Cafiso, S.D.D.D.; Kemnitzer, M.; Guandalini, A.; Kienle, F.; Veronesi, S.; Tonelli, M.; Aus der Au, J.; Agnesi, A. Sub-50-fs widely tunable Yb:CaYAlO4 laser pumped by 400-mW single-mode fiber-coupled laser diode. Opt. Express 2015, 23, 9790. [Google Scholar] [CrossRef]
  73. Wood, D.L.; Nassau, K. Refractive index of cubic zirconia stabilized with yttria. Appl. Opt. 1982, 21, 2978. [Google Scholar] [CrossRef] [PubMed]
  74. Zelmon, D.E.; Small, D.L.; Jundt, D. Infrared corrected Sellmeier coefficients for congruently grown lithium niobate and 5 mol% magnesium oxide –doped lithium niobate. J. Opt. Soc. Am. B 1997, 14, 3319. [Google Scholar] [CrossRef]
  75. Jayasimhadri, M.; Moorthy, L.R.; Saleem, S.A.; Ravikumar, R.V.S.S.N. Spectroscopic characteristics of Sm3+-doped alkali fluorophosphate glasses. Spectrochim. Acta Part A Mol. Biomol. Spectrosc. 2006, 64, 939–944. [Google Scholar] [CrossRef]
  76. Dorenbos, P.; Marsman, M.; Van Eijk, C.W.E.; Korzhik, M.V.; Mlnkov, B.I. Scintillation properties of Y2SiO5:Pr crystals 1. Radiat. Eff. Defects Solids 1995, 135, 325–328. [Google Scholar] [CrossRef]
  77. Jaque, D.; Findensein, J.; Montoya, E.; Capmany, J.; Kaminskii, A.A.; Eichler, H.J.; Solé, J.G. Spectroscopic and laser gain properties of the Nd3+:β′-Gd2(MoO4) 3 non-linear crystal. J. Phys. Condens. Matter 2000, 12, 9699–9714. [Google Scholar] [CrossRef]
  78. Huang, X.; Fang, Q.; Yu, Q.; Lü, X.; Zhang, L.; Lin, Z.; Wang, G. Thermal and polarized spectroscopic characteristics of Nd3+:LiLa(WO4)2 crystal. J. Alloys Compd. 2009, 468, 321–326. [Google Scholar] [CrossRef]
  79. Romanyuk, Y.E.; Borca, C.N.; Pollnau, M.; Rivier, S.; Petrov, V.; Griebner, U. Yb-doped KY(WO4)2 planar waveguide laser. Opt. Lett. 2006, 31, 53–55. [Google Scholar] [CrossRef] [PubMed] [Green Version]
Figure 1. Emission spectra of YF3:Er3+ recorded at 293 K and 473 K, the energy level diagram depicting the emissions 4F7/2, 2H11/2, 4S3/24I15/2 of Er3+, the energy difference between 2H11/2 and 4F7/2 levels from 4S3/2 level (blue and red arrows, respectively), the normalized LIRs of 2H11/24I15/2/4S3/24I15/2 and 4F7/24I15/2/4S3/24I15/2 and the corresponding relative sensitivities on the given temperature range.
Figure 1. Emission spectra of YF3:Er3+ recorded at 293 K and 473 K, the energy level diagram depicting the emissions 4F7/2, 2H11/2, 4S3/24I15/2 of Er3+, the energy difference between 2H11/2 and 4F7/2 levels from 4S3/2 level (blue and red arrows, respectively), the normalized LIRs of 2H11/24I15/2/4S3/24I15/2 and 4F7/24I15/2/4S3/24I15/2 and the corresponding relative sensitivities on the given temperature range.
Crystals 11 00189 g001
Figure 2. Slater integrals and s–o coupling parameters as listed in Table 2.
Figure 2. Slater integrals and s–o coupling parameters as listed in Table 2.
Crystals 11 00189 g002
Figure 3. Judd–Ofelt parameters in different hosts, as given in Table 2.
Figure 3. Judd–Ofelt parameters in different hosts, as given in Table 2.
Crystals 11 00189 g003
Figure 4. Deviation of the RMEs listed in Table 3 from their average values.
Figure 4. Deviation of the RMEs listed in Table 3 from their average values.
Crystals 11 00189 g004
Figure 5. Relative sensitivities vs. absolute sensitivities at 300 K. (ac) for LIR by 2H11/2 higher level, (df) by 4F7/2. (a,d) comparison of crystals and glasses, (b,e) between crystal types, (c,f) between different glasses.
Figure 5. Relative sensitivities vs. absolute sensitivities at 300 K. (ac) for LIR by 2H11/2 higher level, (df) by 4F7/2. (a,d) comparison of crystals and glasses, (b,e) between crystal types, (c,f) between different glasses.
Crystals 11 00189 g005
Figure 6. Relative sensitivities vs. absolute sensitivities at temperatures with maximum absolute sensitivities. (ac) for LIR by 2H11/2 higher level, (df) by 4F7/2. (a,d) comparison of crystals and classes, (b,e) between crystal types, (c,f) between different glasses.
Figure 6. Relative sensitivities vs. absolute sensitivities at temperatures with maximum absolute sensitivities. (ac) for LIR by 2H11/2 higher level, (df) by 4F7/2. (a,d) comparison of crystals and classes, (b,e) between crystal types, (c,f) between different glasses.
Crystals 11 00189 g006
Table 1. A collection of 40 Er3+ doped hosts used in this study. Form: C–Crystal, G–Glass.
Table 1. A collection of 40 Er3+ doped hosts used in this study. Form: C–Crystal, G–Glass.
No.HostAbbreviationNameEr3+ ConcentrationFormRef.
1 *β-NaGdF4 Sodium Gadolinium Fluoride1%C[33]
2LaF3 Lanthanum fluoride0.05%C[34,35]
3Y3Al5O12YAGYttrium Aluminium Garnet1.2 at%C[36]
4LaCl3 Lanthanum Chloride1%C[17]
5La2O2S Lanthanum oxysulfide1 mol%C[37,38]
6Y2O2S Yttrium oxysulfide1 mol%C[37,38]
7 Sb2O3-35P2O5-5MgO-AgClSPMEAAntimony Phosphate0.17 at%G[39]
8YAlO3 Yttrium Orthoaluminate1.5 at%C[40]
9KCaF3 Kalium Calcium Fluoride1.62 at%C[41]
10Y2O3 Yttrium oxide1%C[42,43]
11YVO4 Yttrium Vanadate2.5%C[44]
12PbF2-GaF3-(Zn,Mn)F2PbZnGaLaFLead-Based fluoride0.6%G[45]
13 *CaYAlO4CYAOYttrium Calcium Aluminate0.5 at%C[46]
14Gd3Ga5O12GGGGadolinium Gallium Garnet1.2 at%C[36]
15Y3Sc2Ga3O12YSGGYttrium Scandium Gallium Garnet1.2 at%C[36]
16PbO∙H3BO3∙TiO2∙AlF3LBTAFLead Borate Titanate Aluminium Fluoride4.8 at%G[47]
17Li2CO3∙H3BO3LiBOLithium Borate1 mol%G[48]
18Li2CO3∙H3BO3∙MgCO3MgLiBOMagnesium Lithium Borate1 mol%G[48]
19Li2CO3∙H3BO3∙MgCO3CaLiBOCalcium Lithium Borate1 mol%G[48]
20Li2CO3∙H3BO3∙SrCO3SrLiBOStrontium Lithium Borate1 mol%G[48]
21Li2CO3∙H3BO3∙BaCO3BaLiBOBarium Lithium Borate1 mol%G[48]
22ZrO2∙YO1.5YSZYttria stabilized Zirconia0.7 at%C[49]
23LiNbO3 Lithium Niobate1.5 × 1019 cm−3C[50]
24PbO-PbF2 Oxyfluoride1.35 wt%G[51]
25ZrF4∙BaF2∙LaF3∙AlF3ZBLAFluorozirconate0.5%G[52]
26ZnO∙Al2O3∙Bi2O3∙B2O3ZnAlBiBZinc Alumino Bismuth Borate0.5 mol%G[53]
27NaPO3∙TeO2∙AlF3∙LiFLiTFPLithium Fluorophosphate0.12 at%G[54]
28NaPO3∙TeO2∙AlF3∙NaFNaTFPSodium Fluorophosphate0.12 at%G[54]
29NaPO3∙TeO2∙AlF3∙KFKTFPKalium Fluorophosphate0.12 at%G[54]
30Kigre patented Phosphate1.51 wt%G[55]
31TeO2∙PbF2∙AlF3 Fluoro-tellurite0.625 at%G[56]
32SrGdGa3O7 Strontium Gadolinium Gallium Garnet4.2 × 1021 cm−3C[57]
33YPO4 Yttrium Phosphate0.6 at%C[58]
34Y2SiO5YSOYttrium Orthosilicate2 mol%bC[59]
35BaGd2(MoO4)4BGMBarium Gadolinium Molybdate1.4 at%C[60]
36NaY(MoO4)2NYMSodium Yttrium Molybdate1.25 at%C[61]
37Gd2(MoO4)3 Gadolinium Molybdate1%C[62]
38LiLa(MoO4)2 Lithium Lanthanum Molybdate0.55 at%C[63]
39 *LiLa(WO4)2 Lithium Lanthanum Tungstanate0.65%C[64]
40KY(WO4)2 Kalium Yttrium Tungstanate0.5%C[65]
* Co-doped with Yb3+.
Table 2. Energies of 4S3/2, 2H11/2 and 4F7/2 levels in hosts listed in Table 1, their Slater integrals and spin-orbit (s–o) coupling parameters, and JO intensity parameters as reported in the corresponding references.
Table 2. Energies of 4S3/2, 2H11/2 and 4F7/2 levels in hosts listed in Table 1, their Slater integrals and spin-orbit (s–o) coupling parameters, and JO intensity parameters as reported in the corresponding references.
No.HostEnergy [cm−1]Slater Integrals and s–o [cm−1]JO Parameters × 1020 [cm2]
4S3/22H11/24F7/2F2F-4F6ζΩ2Ω4Ω6
1β–NaGdF4184591918620483429.468.77.124034.971.162.03
2LaF3183531911820412435.767.27.423513.91.02.3
3YAG181661896720344433.265.06.523450.740.920.70
4LaCl3183831906820414433.266.97.323865.452.080.69
5La2O2S182361893020320445.866.67.523954.322.321.17
6Y2O2S182361893020320444.068.17.423912.712.101.85
7 dSPMEA181821917520576NANANANA7.893.271.07
8YAlO3183501915020300NANANANA0.950.580.55
9 dKCaF3184501930520576NANANANA0.740.870.57
10 bY2O3180711893120266429.665.07.123834.591.210.48
11 bYVO4182091905920371440.866.87.3238113.452.231.67
12PbZnGaLaF185521919320618444.064.66.923951.541.131.19
13CYAO182981904020454NANANANA3.782.521.91
14 cGGG184501916820387NANANANA0.700.370.86
15 cYSGG184331912720345NANANANA0.920.480.87
16 aLBTAF183821919420450433.967.06.723865.891.101.47
17 aLiBO184131920520488437.865.16.823773.240.920.82
18 aMgLiBO183891916820462437.864.56.823731.330.390.62
19 aCaLiBO184131920520488438.264.96.823763.680.761.52
20 aSrLiBO184131920520488438.664.76.823812.530.391.10
21 aBaLiBO184031920520488438.264.96.823821.800.280.90
22 cYSZ184161934220534NANANANA1.500.500.22
23 cLiNbO3182481904720492NANANANA7.292.241.27
24 aPbO-PbF2185011929720601437.866.67.324613.221.340.61
25ZBLA184501919320534NANANANA2.541.390.97
26 cZnAlBiB184841919320491NANANANA2.101.531.43
27LiTFP183441918920486NANANANA4.701.211.30
28NaTFP183771922620486NANANANA5.921.071.44
29KTFP183771922620486NANANANA5.090.691.45
30Kigre patented183501915020300NANANANA6.281.031.39
31 cTeO2∙PbF2∙AlF3183321895720434NANANANA5.522.071.00
32SrGdGa3O7181351913120411NANANANA2.461.240.51
33YPO4183481908320449435.767.27.423513.023.072.58
34 eYSO183481908320449NANANANA1.290.292.78
35 cBGM183481908320408NANANANA12.331.960.96
36 cNYM181481915720449NANANANA13.341.692.29
37Gd2(MoO4)3183481915720449NANANANA11.748.163.98
38LiLa(MoO4)2183481915720449NANANANA8.071.060.83
39LiLa(WO4)2183481904720345NANANANA9.032.020.59
40KY(WO4)2184201919020450NANANANA7.082.301.01
a Slater integrals calculated from F2,4,6 parameters by Equation (15) in Ref. [17]. b Slater integrals calculated from Racah parameters by Equation (17) in Ref. [17]. c Slater integrals and spin–orbit coupling parameter not provided, the RME values the authors used are by Carnall in Ref. [66], or by d Weber in Ref. [35]. e Energy levels are not given in the literature, values in the table are provided approximately.
Table 3. Squared RMEs for hosts in Table 1, recalculated by RELIC software [17] from Slater integrals and s–o coupling parameters in Table 2, refractive index values, local corrections for emission, and induced electric dipole strengths. Note: if Slater integrals were not provided in Table 2, the squared RMEs will be given from the tables by Carnall [66], unless indicated that the authors used tables by Weber [35].
Table 3. Squared RMEs for hosts in Table 1, recalculated by RELIC software [17] from Slater integrals and s–o coupling parameters in Table 2, refractive index values, local corrections for emission, and induced electric dipole strengths. Note: if Slater integrals were not provided in Table 2, the squared RMEs will be given from the tables by Carnall [66], unless indicated that the authors used tables by Weber [35].
No.Initial LevelSquared RMEnχED (Emission)Ref. for nD [esu2 cm2] × 1040
U2U4U6
1 b4S3/2000.22161.4993.00[33]10.38
2H11/20.72470.41590.09251.4993.0098.55
4F7/200.14610.62981.4993.0033.40
24S3/2000.22751.5163.11[67]12.07
2H11/20.71410.41120.08671.5183.1278.33
4F7/200.14730.62851.5223.1536.75
34S3/2000.21341.8365.88[36]3.45
2H11/20.58160.33500.07561.8385.9118.26
4F7/200.14650.61921.8425.9513.11
4 b4S3/2000.22261.74.52[68]3.54
2H11/20.72050.41520.09111.74.52111.96
4F7/200.14670.62741.74.5217.03
5 b4S3/2000.22402.211.44[69]6.05
2H11/20.68720.39710.08492.211.4492.03
4F7/200.14740.62472.211.4424.75
6 b4S3/2000.22572.211.44[69]9.63
2H11/20.68910.39680.08412.211.4465.89
4F7/200.14720.62722.211.4433.90
7 c4S3/2000.22112.3514.70[39]5.64
2H11/20.71250.41230.09252.3514.70161.57
4F7/200.14680.62662.3514.7026.55
8 a4S3/2000.22111.9467.24[40,41]2.81
2H11/20.71250.41230.09251.9487.2722.30
4F7/200.14680.62661.9537.349.91
94S3/2000.22851.4022.45[35]3.00
2H11/20.70560.41090.08701.4042.4621.44
4F7/200.14670.62731.4062.4711.19
104S3/2000.21711.9387.13[70]2.40
2H11/20.69640.40220.09121.9427.1985.98
4F7/200.14640.62381.9487.2710.99
114S3/2000.22312.0178.25[71]8.59
2H11/20.67960.39190.08432.0238.34234.27
4F7/200.14710.62532.0368.5431.66
12 b4S3/2000.21821.6113.78[45]5.99
2H11/20.65470.37950.08381.6113.7835.45
4F7/200.14710.62001.6113.7820.85
13 bd4S3/2000.22851.856.04[72]10.07
2H11/20.70560.41090.08701.856.0489.25
4F7/200.14670.62731.856.0436.17
14 ac4S3/2000.22111.9877.81[36]4.39
2H11/20.71250.41230.09251.9827.7416.86
4F7/200.14680.62661.9987.9713.68
15 ac4S3/2000.22111.9447.21[36]4.44
2H11/20.71250.41230.09251.9487.2721.54
4F7/200.14680.62661.9547.3514.20
16 b4S3/2000.21571.5643.44[47]7.31
2H11/20.62960.36180.08151.5643.4497.49
4F7/200.14630.62351.5643.4424.86
17 b4S3/2000.21451.4782.88[48]4.06
2H11/20.61150.35310.07951.4782.8854.70
4F7/200.14660.61961.4782.8814.83
18 b4S3/2000.21381.4762.86[48]3.06
2H11/20.60680.35070.07921.4762.8622.91
4F7/200.14670.61851.4762.8610.17
19 b4S3/2000.21431.4802.89[48]7.51
2H11/20.60810.35120.07911.4802.8960.55
4F7/200.14670.61911.4802.8924.28
20 b4S3/2000.21351.4792.88[48]5.42
2H11/20.60670.35070.07941.4792.8840.58
4F7/200.14660.61841.4792.8817.01
21 b4S3/2000.21371.4812.89[48]4.44
2H11/20.61000.35240.07981.4812.8929.26
4F7/200.14660.61901.4812.8913.80
22 c4S3/2000.22112.16710.80[73]1.12
2H11/20.71250.41230.09252.17210.8929.88
4F7/200.14680.62662.18011.044.87
23 c4S3/2000.22112.31613.95[74]6.48
2H11/20.71250.41230.09252.33114.31143.84
4F7/200.14680.62662.34914.7525.94
24 b4S3/2000.21561.7795.27[51]3.03
2H11/20.71760.41470.09591.7795.2767.47
4F7/200.14610.62431.7795.2713.30
25 d4S3/2000.22851.5183.12[52]5.11
2H11/20.70560.41090.08701.5193.1356.47
4F7/200.14670.62731.5203.1418.74
26 b4S3/2000.22111.8195.70[53]7.29
2H11/20.71250.41230.09251.8195.7052.12
4F7/200.14680.62661.8195.7025.85
27 bd4S3/2000.22851.5843.58[75]6.85
2H11/20.70560.41090.08701.5843.5890.58
4F7/200.14670.62731.5843.5822.91
28 bd4S3/2000.22851.5873.60[75]7.59
2H11/20.70560.41090.08701.5873.60109.39
4F7/200.14670.62731.5873.6024.46
29 bd4S3/2000.22851.5883.61[75]7.64
2H11/20.70560.41090.08701.5883.6192.30
4F7/200.14670.62731.5883.6123.32
30 a4S3/2000.22111.5813.56[55]7.09
2H11/20.71250.41250.09251.5873.60115.99
4F7/200.14690.62661.5993.6923.58
31 bc4S3/2000.22112.1169.86[56]5.10
2H11/20.71250.41230.09252.1169.86112.55
4F7/200.14680.62662.1169.8621.46
32 bd4S3/2000.22851.8315.83[57]2.69
2H11/20.70560.41090.08701.8315.8352.82
4F7/200.14670.62731.8315.8311.58
33 b4S3/2000.22751.775.18[58]13.54
2H11/20.71410.41120.08661.775.1884.03
4F7/200.14730.62851.775.1847.84
34 bd4S3/2000.22851.85.49[76]14.65
2H11/20.70560.41090.08701.85.4929.33
4F7/200.14670.62731.85.4941.21
35 bc4S3/2000.22112.028.30[60]4.90
2H11/20.71250.41230.09252.028.30223.35
4F7/200.14680.62662.028.3020.51
36 c4S3/2000.22112.018.15[61]11.68
2H11/20.71250.41230.09252.008.00240.22
4F7/200.14680.62662.008.0038.82
37 ad4S3/2000.22112.1610.66[55,77]20.30
2H11/20.71250.41250.09252.1610.66279.11
4F7/200.14690.62662.1610.6685.18
38 d4S3/2000.22852.058.76[63]4.38
2H11/20.70560.41090.08702.058.76143.07
4F7/200.14670.62732.058.7615.60
39 bd4S3/2000.22852.08.00[78]3.11
2H11/20.70560.41090.08702.08.00167.32
4F7/200.14670.62732.08.0015.37
40 bd4S3/2000.22852.08.00[79]5.32
2H11/20.70560.41090.08702.08.00139.07
4F7/200.14670.62732.08.0022.40
a RME values not calculated by RELIC software, but given in the corresponding reference. b Refractive Index values approx. wavelength-independent. c RME values from Carnall [66], d from Weber [35].
Table 4. Average RME values estimated from squared RMEs listed in Table 3. Deviations of average values from squared RMEs reported by Carnall (C) and Weber (W), in percentage.
Table 4. Average RME values estimated from squared RMEs listed in Table 3. Deviations of average values from squared RMEs reported by Carnall (C) and Weber (W), in percentage.
Initial LevelU2U4U6U2(C)U4(C)U6(C)U2(W)U4(W)U6(W)
4S3/2000.2224000.6002.7
2H11/20.68890.39890.08743.33.35.52.42.90.5
4F7/200.14680.6254000.200.10.3
Table 5. Calculated luminescence thermometry parameters: energy gaps (ΔE) from Er3+ 4S3/2 level to 2H11/2 and 4F7/2, relative temperature sensitivities (Sr) for LIRs between selected levels, B LIR parameters, absolute sensitivities at room temperature (Sa), maximum sensitivity value (Samax), temperatures at which maximum absolute sensitivity occurs (T(Samax)), and relative sensitivities at T(Samax) (Sr(T(Samax)).
Table 5. Calculated luminescence thermometry parameters: energy gaps (ΔE) from Er3+ 4S3/2 level to 2H11/2 and 4F7/2, relative temperature sensitivities (Sr) for LIRs between selected levels, B LIR parameters, absolute sensitivities at room temperature (Sa), maximum sensitivity value (Samax), temperatures at which maximum absolute sensitivity occurs (T(Samax)), and relative sensitivities at T(Samax) (Sr(T(Samax)).
No.Higher LevelΔESr (300 K) [% K−1]BSa (300 K) [K−1]max(Sa) [K−1]T(max(Sa)) [K] (°C)Sr (T(max(Sa)) [% K−1]
12H11/27271.1610.660.0037920.0055523 (250)0.38
4F7/220243.244.400.0000090.00081456 (1183)0.14
22H11/27651.227.370.0022970.0036550 (277)0.36
4F7/220593.294.240.0000070.00081481 (1208)0.14
32H11/28011.286.050.0016630.0028576 (303)0.35
4F7/221783.485.400.0000050.00091567 (1294)0.13
42H11/26851.1035.260.0144530.0194493 (220)0.41
4F7/220313.256.580.0000130.00121461 (1188)0.14
52H11/26941.1017.030.0067720.0092499 (226)0.40
4F7/220843.305.660.0000090.00101499 (1226)0.13
62H11/26941.107.650.0030430.0042499 (226)0.40
4F7/220843.304.870.0000070.00091499 (1226)0.13
72H11/29931.5933.600.0045570.0128714 (441)0.28
4F7/223943.836.820.0000030.00111722 (1449)0.12
82H11/28001.289.070.0025010.0043576 (303)0.35
4F7/219503.124.850.0000130.00091403 (1130)0.14
92H11/28551.378.210.0018580.0036615 (342)0.33
4F7/221263.405.210.0000070.00091529 (1256)0.13
102H11/28601.3741.420.0092090.0181619 (346)0.32
4F7/221953.516.570.0000060.00111579 (1306)0.13
112H11/28501.3631.600.0072840.0140612 (339)0.33
4F7/221623.465.340.0000060.00091555 (1282)0.13
122H11/26411.026.550.0031040.0038461 (188)0.43
4F7/220663.304.780.0000080.00091486 (1213)0.13
132H11/27421.199.990.0033730.0051534 (261)0.37
4F7/221563.455.020.0000060.00091551 (1278)0.13
142H11/27181.154.270.0015660.0022517 (244)0.39
4F7/219373.104.290.0000120.00081394 (1121)0.14
152H11/26941.115.460.0021730.0030499 (226)0.40
4F7/219123.064.380.0000140.00091376 (1103)0.15
162H11/28121.3015.170.0040090.0070584 (311)0.34
4F7/220683.314.680.0000080.00091488 (1215)0.13
172H11/27921.2715.300.0043390.0073570 (297)0.35
4F7/220753.325.040.0000080.00091493 (1220)0.13
182H11/27791.258.480.0025190.0041560 (287)0.36
4F7/220733.314.580.0000070.00081491 (1218)0.13
192H11/27921.279.140.0025940.0043570 (297)0.35
4F7/220753.324.450.0000070.00081493 (1220)0.13
202H11/27921.278.500.0024110.0040570 (297)0.35
4F7/220753.324.330.0000070.00081493 (1220)0.13
212H11/28021.287.500.0020520.0035577 (304)0.35
4F7/220853.334.290.0000060.00081500 (1227)0.13
222H11/29261.4831.120.0054280.0126666 (393)0.30
4F7/221183.396.160.0000080.00111524 (1251)0.13
232H11/27991.2825.900.0071670.0122575 (302)0.35
4F7/222443.596.000.0000050.00101614 (1341)0.12
242H11/27961.2725.240.0070580.0119573 (300)0.35
4F7/221003.366.050.0000090.00111511 (1238)0.13
252H11/27431.1912.460.0041940.0063535 (262)0.37
4F7/220843.335.070.0000080.00091499 (1226)0.13
262H11/27091.138.000.0030250.0042510 (237)0.39
4F7/220073.214.830.0000100.00091444 (1171)0.14
272H11/28451.3515.130.0035510.0067608 (335)0.33
4F7/221423.424.660.0000060.00081541 (1268)0.13
282H11/28491.3616.500.0038180.0073611 (338)0.33
4F7/221093.374.460.0000060.00081517 (1244)0.13
292H11/28491.3613.830.0031990.0061611 (338)0.33
4F7/221093.374.230.0000060.00081517 (1244)0.13
302H11/28001.2818.820.0051900.0089576 (303)0.35
4F7/219503.124.670.0000130.00091403 (1130)0.14
312H11/26251.0024.400.0121680.0147450 (177)0.44
4F7/221023.365.830.0000080.00101512 (1239)0.13
322H11/29961.5923.070.0030930.0087717 (444)0.28
4F7/222763.646.140.0000040.00101637 (1364)0.12
332H11/27351.186.980.0024160.0036529 (256)0.38
4F7/221013.364.890.0000070.00091512 (1239)0.13
342H11/27351.182.250.0007790.0012529 (256)0.38
4F7/221013.363.890.0000050.00071512 (1239)0.13
352H11/27351.1851.320.0177570.0263529 (256)0.38
4F7/220603.295.770.0000100.00111482 (1209)0.13
362H11/210091.6123.750.0030320.0089726 (453)0.28
4F7/223013.684.670.0000030.00081655 (1382)0.12
372H11/28091.2915.650.0041790.0073582 (309)0.34
4F7/221013.365.810.0000080.00101512 (1239)0.13
382H11/28091.2937.220.0099400.0173582 (309)0.34
4F7/221013.364.940.0000070.00091512 (1239)0.13
392H11/26991.1260.180.0235370.0324503 (230)0.40
4F7/219973.196.740.0000150.00131437 (1164)0.14
402H11/27701.2329.540.0090520.0144554 (281)0.36
4F7/220303.255.760.0000110.00111460 (1187)0.14
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Ćirić, A.; Gavrilović, T.; Dramićanin, M.D. Luminescence Intensity Ratio Thermometry with Er3+: Performance Overview. Crystals 2021, 11, 189. https://doi.org/10.3390/cryst11020189

AMA Style

Ćirić A, Gavrilović T, Dramićanin MD. Luminescence Intensity Ratio Thermometry with Er3+: Performance Overview. Crystals. 2021; 11(2):189. https://doi.org/10.3390/cryst11020189

Chicago/Turabian Style

Ćirić, Aleksandar, Tamara Gavrilović, and Miroslav D. Dramićanin. 2021. "Luminescence Intensity Ratio Thermometry with Er3+: Performance Overview" Crystals 11, no. 2: 189. https://doi.org/10.3390/cryst11020189

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