Next Article in Journal
Globular Flower-Like Reduced Graphene Oxide Design for Enhancing Thermally Conductive Properties of Silicone-Based Spherical Alumina Composites
Next Article in Special Issue
Comparison of Three Ratiometric Temperature Readings from the Er3+ Upconversion Emission
Previous Article in Journal
Green Silver Nanoparticles Formed by Phyllanthus urinaria, Pouzolzia zeylanica, and Scoparia dulcis Leaf Extracts and the Antifungal Activity
Previous Article in Special Issue
Synthesis, Cytotoxicity Assessment and Optical Properties Characterization of Colloidal GdPO4:Mn2+, Eu3+ for High Sensitivity Luminescent Nanothermometers Operating in the Physiological Temperature Range
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Making Nd3+ a Sensitive Luminescent Thermometer for Physiological Temperatures—An Account of Pitfalls in Boltzmann Thermometry

1
Condensed Matter and Interfaces, Debye Institute for Nanomaterials Science, Department of Chemistry, Utrecht University, Princetonplein 1, 3584 CC Utrecht, The Netherlands
2
Vinča Institute of Nuclear Sciences, University of Belgrade, 11001 Belgrade, Serbia
*
Author to whom correspondence should be addressed.
Submission received: 22 February 2020 / Revised: 12 March 2020 / Accepted: 13 March 2020 / Published: 18 March 2020
(This article belongs to the Special Issue Luminescent Rare-Earth-Based Nanomaterials)

Abstract

:
Ratiometric luminescence thermometry employing luminescence within the biological transparency windows provides high potential for biothermal imaging. Nd3+ is a promising candidate for that purpose due to its intense radiative transitions within biological windows (BWs) I and II and the simultaneous efficient excitability within BW I. This makes Nd3+ almost unique among all lanthanides. Typically, emission from the two 4F3/2 crystal field levels is used for thermometry but the small ~100 cm−1 energy separation limits the sensitivity. A higher sensitivity for physiological temperatures is possible using the luminescence intensity ratio (LIR) of the emissive transitions from the 4F5/2 and 4F3/2 excited spin-orbit levels. Herein, we demonstrate and discuss various pitfalls that can occur in Boltzmann thermometry if this particular LIR is used for physiological temperature sensing. Both microcrystalline, dilute (0.1%) Nd3+-doped LaPO4 and LaPO4: x% Nd3+ (x = 2, 5, 10, 25, 100) nanocrystals serve as an illustrative example. Besides structural and optical characterization of those luminescent thermometers, the impact and consequences of the Nd3+ concentration on their luminescence and performance as Boltzmann-based thermometers are analyzed. For low Nd3+ concentrations, Boltzmann equilibrium starts just around 300 K. At higher Nd3+ concentrations, cross-relaxation processes enhance the decay rates of the 4F3/2 and 4F5/2 levels making the decay faster than the equilibration rates between the levels. It is shown that the onset of the useful temperature sensing range shifts to higher temperatures, even above ~ 450 K for Nd concentrations over 5%. A microscopic explanation for pitfalls in Boltzmann thermometry with Nd3+ is finally given and guidelines for the usability of this lanthanide ion in the field of physiological temperature sensing are elaborated. Insight in competition between thermal coupling through non-radiative transitions and population decay through cross-relaxation of the 4F5/2 and 4F3/2 spin-orbit levels of Nd3+ makes it possible to tailor the thermometric performance of Nd3+ to enable physiological temperature sensing.

Graphical Abstract

1. Introduction

Temperature is an important control parameter that governs, e.g., the rate of chemical reactions, but also the optimum working efficiency of electronic devices or dynamics and viability of biological systems. As such, non-invasive, sensitive and remote temperature measurement techniques with the capability to spatially resolve temperature variations down to the micrometer range are becoming increasingly relevant [1,2,3,4,5,6,7]. Physiological temperature sensing is especially demanding in that regard as it even requires to accurately distinguish temperature fluctuations below 1 K. Luminescence nanothermometry is an appealing and rapidly emerging technique that meets those requirements and constantly improves [8,9,10,11,12,13,14]. Among the various temperature-dependent optical parameters that can be used for temperature calibration, the luminescence intensity ratio (LIR) of two emissive transitions from thermally coupled excited states has emerged as an especially robust and yet easily measurable representative to extract information about the local temperature of a medium in contact with luminescent nanocrystals. Also, other methodological approaches such as lifetime thermometry [15,16,17,18,19,20,21], or the concept of excited state absorption (ESA) [22,23,24] have been developed for that purpose while organic framework-based thermometers increasingly attract attention [25,26].
Besides the requirement for a low temperature uncertainty (<±1 K) over a narrow temperature range (30 °C–60 °C), luminescence nanothermometry under physiological conditions has additional restrictions. Absorbance, light scattering or even autofluorescence of locally surrounding tissue pose challenges that have to be overcome to guarantee a reliable detection of the LIR outside the biological system of interest [27,28,29]. Thus, near infrared (NIR) luminescence within the biological transparency windows (BW I: 650–950 nm; BW II: 1000–1350 nm; BW III: 1550–1850 nm) is highly desirable for minimized attenuation of the emitted light to guarantee high temperature precision. While quantum dots rely on a sensitive thermal quenching behavior of their NIR luminescence within physiological temperature regimes and have successfully been employed in various biomedical applications [30,31,32,33,34,35,36], their usage in thermometry still requires careful calibration [37].
As an alternative, the lanthanide ion Nd3+(4f3) has become a promising candidate for in vivo nanothermometry [30,38,39,40,41,42,43]. This is related to the fact that its most intense radiative transitions all lie within BW I and BW II. Moreover, the large energy gap between the most dominantly emissive 4F3/2 spin-orbit level and the next lower 4I15/2 level typically allows for rather high brightness of Nd3+-activated luminescent nanocrystals. Finally, Nd3+ most efficiently absorbs within the first biological window. Altogether, those optical properties make Nd3+ almost unique among all lanthanides for NIR luminescence thermometry. Since the 4F3/2 spin-orbit level of Nd3+ splits into exactly two energetically well isolated Kramers’ doublets (R1 and R2) in any lower symmetric crystal field with an energy gap of around 100 cm−1, the vast majority of studies aiming at single ion in vivo nanothermometry with Nd3+ addressed the LIR stemming from emission from those two crystal field states [38,40,41,44,45]. Among the various possibilities of synthetically well accessible and size-controllable nanocrystals such as fluorides [46,47,48], nanosized garnets have become especially interesting in that sense due to the well resolved and intense 4F3/2(R1,2) → 4I9/2(Z5) emissive transitions at 935 nm and 945 nm. This allowed researchers to tune the relative thermal sensitivity Sr of this particular LIR towards the theoretically expected maximum value of around 0.15% K−1 at 37 °C according to the given energy gap [49,50]. The value obtained with that concept of Nd3+-based nanothermometry could only be exceeded in lattices with higher crystal field splitting, for example LiLuF4 nanocrystals, with a value of around 0.5% K−1 at 37 °C [51].
Usage of the two Kramers’ doublets for in vivo nanothermometry is, however, connected to some inherent problems. Besides the challenges in size and morphology control of nanosized garnets [52] compared to, e.g., well-established synthesis routes to fluoride nanocrystals [46,47,48], the requirement for high spectral resolution and unambiguous assignment of the two closely lying emissive transitions to be attained under real in vivo conditions such as light scattering and absorption of surrounding tissue, dynamic flow, or movement of biological entities, is challenging. Even more severely, the theoretically achievable maximum relative sensitivity is well below 1% K−1 and does not allow to achieve the required temperature measurement precision below 1 K even with use of laser excitation sources and sensitive light detection systems. Finally, the small energy gap between the two Kramers’ doublets inherently diminishes the absolute thermal response of the LIR because the two states have similar population densities at room temperature, as has been shown explicitly by thermodynamic modelling of a generalized excited two-level system to be reported in detail elsewhere [53].
One way to overcome this problem and achieve an order of magnitude higher relative sensitivities has already been demonstrated by a phonon-assisted energy transfer between Nd3+ and Yb3+ [30,54,55] or recently, also Nd3+ and Er3+ [43]. However, the underlying temperature calibration model for the temperature-dependent LIR based on energy transfer is far from trivial and, strictly stated, Boltzmann’s law does not apply for those situations [53]. In addition, the NIR emission of Yb3+ lies in between the two transparency windows BW I and II and strongly overlaps with the 4F3/24I11/2-related emission of Nd3+ thus introducing additional error sources for accurate ratiometric luminescence thermometry.
A promising alternative that would allow for calibration with Boltzmann’s law and does not require the introduction of additional dopants is usage of the two spin-orbit levels 4F5/2 and 4F3/2 of Nd3+ instead. They have an effective energy separation of roughly 1000 cm−1 and emission from those two levels can be easily spectrally resolved. In addition, the higher energy gap permits an order of magnitude higher relative sensitivity (Sr > 1% K−1) at physiological temperatures, which meets the requirements for in vivo nanothermometry. Besides an original work by Haro-González et al. in a Nd3+-doped Sr-Ba-based niobate glass demonstrating a proof of principle [56], a more elaborate analysis of thermometry with that energy gap has been recently reported by Kolesnikov et al. in Y2O3:Nd3+ nanocrystals [57]. However, thermal coupling between the 4F5/2 and 4F3/2 spin-orbit levels is an issue because of the relatively large gap of around 1000 cm−1. The non-radiative transition rates between the two spin-orbit levels have to be faster than decay rates of these levels within the relevant temperature regime since otherwise Boltzmann-based equilibrium is not sustained [58]. In addition, higher doping concentrations to help increase the absorption efficiency of Nd3+ may also induce faster decay as a result of energy migration and cross-relaxation effects, preventing Boltzmann equilibrium. Thus, dependent on the electronic structure of a lanthanide ion, the effect of energy migration or cross relaxation at higher doping concentrations on the Boltzmann equilibrium between two excited levels of interest has to be investigated in detail.
Here we present and discuss the 4F5/2 and 4F3/2 based alternative thermometric sensing by Nd3+ and evaluate its applicability for in vivo nanothermometry using LaPO4:Nd3+ nanocrystals as an illustrative example. The choice of the host material arises from the fact that phosphates have a large maximum phonon energy of around 1050 cm−1 that allows for resonant thermalization between the 4F5/2 and 4F3/2 levels [59]. Moreover, phosphate nanocrystals may also be at least theoretically expected to show biocompatibility given its presence to a large fraction in the body of mammals. Besides a structural and optical characterization, especially the impact of both temperature and Nd3+ concentration on the thermometric performance of those nanocrystals will be analyzed. We will carefully compare the found results to the luminescence decay kinetics of the Nd3+ ions and compare predictions with experimental results. Our general purpose is to draw attention to some potential pitfalls in conventional Boltzmann thermometers, discuss their origin and provide solutions.

2. Materials and Methods

2.1. Synthesis

La1−xNdxPO4 nanocrystals (x = 0.02, 0.05, 0.10, 0.25, 1.00) were synthesized with a common hydrothermal co-precipitation approach. For that, stoichiometric amounts of La(NO3)3 · 6 H2O (Alfa Aesar, 99.99%, Karlsruhe, Germany) and Nd(NO3)3 · 6 H2O (Alfa Aesar, 99.9%, Karlsruhe, Germany) were dissolved in deionized water. A small volume of 1 M NaOH (Chemos, 99%, Altdorf, Germany) solution was added under stirring of the solution to attain an alkaline medium. Afterwards, an aqueous solution containing a stoichiometric amount of (NH4)2HPO4 (Alfa Aesar, 98%, Karlsruhe, Germany) was added to allow for precipitation of the rare earth phosphates. The resulting solution was transferred to a Teflon-lined stainless-steel autoclave and kept at 140 °C for 30 h to induce crystallization of the nanoseeds of the rare earth phosphates. After natural cooling to room temperature, the colorless precipitate was collected, washed with deionized H2O and ethanol (EtOH) several times and finally dried at 40 °C for 22 h.
Microcrystalline, dilute La0.999Nd0.001PO4 powder was synthesized for control experiments and to elucidate concentration effects. The synthesis was similarly performed by means of a co-precipitation approach from dissolved La(NO3)3 · 6 H2O (Sigma Aldrich, 99.99%, Schnelldorf, Germany) and Nd(NO3)3 · 6 H2O (Sigma Aldrich, 99.9%, Schnelldorf, Germany) in alkaline aqueous solution and precipitation with a saturated (NH4)2HPO4 (Sigma Aldrich, 99%, Schnelldorf, Germany) solution. The colorless precipitated phosphate was filtered off, washed with deionized H2O three times and carefully dried at 150 °C on air for 3 h. After grinding of the dried residue to a fine powder, it was finally annealed at 1000 °C in air for 3 days.

2.2. Structural and Morphological Characterization

The La1−xNdxPO4 (x = 0.02, 0.05, 0.10, 0.25, 1.00) nanocrystalline powders were characterized by means of X-ray powder diffraction (XRPD) and transmission electron microscopy (TEM). XRPD patterns were measured on a SmartLab (Rigaku, Tokyo, Japan) instrument using Cu Kα radiation (λ = 1.54056 Å). The average crystallite size and internal strain of the nanocrystals were determined by a PDXL2 built-in package software using the known crystal structures of LaPO4 and NdPO4 as a structural input. The XRPD pattern of La0.999Nd0.001PO4 was measured on a Philips PW391 X-ray diffractometer (Eindhoven, The Netherlands) with Cu Kα radiation (λ = 1.54056 Å) in reflection mode with an Al sample holder.
TEM images were acquired on a Thermo Fisher Scientific (formerly Philips, Eindhoven, The Netherlands) FEI Tecnai 12 microscope with an electron acceleration voltage of U = 100 kV. Samples were prepared by scooping a Cu TEM grid with carbon-coated polymer support film into the nanopowders.

2.3. Diffuse Powder Reflectance Spectroscopy

Diffuse reflectance spectra were acquired on a Shimadzu (Kyoto, Japan) UV-Visible UV-2600 spectrophotometer equipped with an integrated sphere (ISR-2600 Plus). All diffuse reflectance spectra were corrected for background using a BaSO4 standard in the regarded spectral range (200–1200 nm).

2.4. (Time-Resolved) Luminescence Spectroscopy and Thermometry

Luminescence spectra were measured on an Edinburgh FLS920 spectrofluorometer (Livingston, UK) equipped with 0.25 m single grating monochromators and detected with a Hamamatsu R5509-72 NIR photomultiplier tube (PMT, Hamamatsu, Shizuoka, Japan) that was cooled with liquid N2. All emission spectra were acquired by excitation with continuous wave (CW) power-tunable (power limit: Pmax = 800 mW, Livingston, UK) lasers operating at λex = 690 nm. A high-resolution emission spectrum to evaluate the impact of non-radiative relaxation from the 4F5/2 to the 4F3/2 level was measured upon excitation with a CW power-tunable (power limit: Pmax = 2 W, Livingston, UK) laser with λex = 808 nm. The incident laser power in each experiment was set to a sufficiently low value in order to avoid laser-induced heating with yet high signal-to-noise ratio. Photoluminescence excitation spectra were acquired upon monitoring the most intense emission of Nd3+ at λem = 1057 nm and a 450 W Xe lamp as external excitation source. Luminescence decay curves in the microsecond domain were obtained by external excitation with a pulsed wavelength-tunable Opotek (Carlsbad, CA, USA) Opolette 355 LD optical parametric oscillator (OPO) with a repetition rate of 20 Hz and temporal pulse width of around 6 ns. The time-resolved signal was detected with a multichannel scaler (MCS) connected to the NIR PMT. Temperature-dependent emission spectra above room temperature were measured in a Linkam (Surrey, UK) THMS600 Microscope Stage (±0.1 °C temperature stability) that could be placed in the spectrometer.

3. Results and Discussion

3.1. Structural and Morphological Characterization of the Nd3+-Activated LaPO4 Nanocrystals

Both LaPO4 and NdPO4 crystallize in a monazite structure type with monoclinic crystal system and space group P21/n (no. 14) [60,61]. Only one crystallographically independent La3+ or Nd3+ site on the Wyckoff position 4e are present in both unit cells, respectively. Given the only slightly smaller ionic radius of Nd3+ (1.163 Å for nine-fold coordination) compared to that of La3+ (1.216 Å for nine-fold coordination) [62], a full range of solubility of NdPO4 within the LaPO4 host is expected. Synthesis of the solid solutions La1−xNdxPO4 by a co-precipitation approach instead conventional heterogeneous mixing and subsequent thermal annealing permits a more random distribution of the Nd3+ activators substituting for the La3+ ions, as has also been recently independently established by solid state magic angle spinning nuclear magnetic resonance (MAS-NMR) experiments on the 31P nuclei in lanthanide-activated LaPO4, whose resonances sensitively react on paramagnetic impurities in their close environment [63,64,65].
Figure 1a depicts the XRPD patterns of the Nd3+-activated LaPO4 nanocrystals as prepared by the hydrothermal co-precipitation approach. The broad background in the low 2θ regime as well as the low intensities and large widths of most Bragg reflections already indicate a small average crystallite size of the particles. Rietveld refinement of the X-ray diffraction patterns employing the monazite structure type as structural input [60,61], affords estimated average crystallite sizes between 6 nm and 10 nm (see Table 1). The small R factors below 10% and the goodness of fit (G.o.f.) parameter close to 1 indicate a very good agreement between the theoretically expected powder diffraction pattern according to the structural input of the monazite-type phases and the experimentally measured diffraction patterns (see Table 1). Moreover, the resulting strain percentage within the nanocrystals is close to 0%, which reflects the miscibility of the two constituents LaPO4 and NdPO4 in the solid solution.
Additional evidence for the homogeneous distribution of the Nd3+ ions substituting for the La3+ sites in LaPO4 is given by observation of a gradual shift of the Bragg reflections towards higher values of 2θ with increasing Nd3+ content in the nanocrystals (see Figure 1b). The microcrystalline sample was separated from that analysis since the well-defined reflections in that particular case merge together upon particle-size induced broadening of the Bragg reflections in the nanocrystals.
This effect leads to artificial shifts to yet lower Bragg angles despite an increase of the Nd3+ content from x = 0.001 to x = 0.02. Accordingly, the resulting cell volume as obtained from the Rietveld refinement also gradually decreases with higher Nd3+ content (see Table 1).
The TEM images (see Figure 2) of the La1−xNdxPO4 nanocrystals confirm the estimated crystallite sizes according to the XRPD patterns and diameters in the range of 10 nm to 15 nm are found. The nanocrystals have an anisotropic rod-like shape, which is also expected given the monoclinic crystal system. An average aspect ratio of around length/width = 1.5 is found for all nanocrystals. Moreover, the TEM images reveal a strong aggregation tendency of the nanocrystals in the powder, which was also found previously in higher condensed La-based phosphates such as the tetraphosphates ALa1−yNdy(PO3)4 (A = Li – Rb; 0 ≤ y ≤ 1) [66].

3.2. Diffuse Reflectance and Optical Absorption

Diffuse reflectance spectra can serve as an additional characterization tool for the successful incorporation of Nd3+ ions into the synthesized Nd3+-activated LaPO4 nanocrystals. The very low Nd3+ concentration in microcrystalline La0.999Nd0.0001PO4 did not give rise to a measurable reflectance signal and was thus excluded. A measure for the absorption coefficient of the respective powdered samples is accessible by the Kubelka-Munk function K/S under the assumption of a constant scattering part of the powders:
K S = f ( R ) = ( 1     R ) 2 2 R ,
where R denotes the diffuse reflectance in the limit of a much higher scale of layer thickness of the powder compared to the average crystallite size, which is clearly fulfilled under the employed measurement conditions. The Kubelka-Munk spectra are depicted in Figure 3a. As expected, the narrow 4f3(4I9/2) → 4f3(2S+1LJ) transitions of Nd3+ show an increasing absorption strength with increasing Nd3+ content of the nanocrystals. Indeed, a linear correlation between the integrated Kubelka-Munk signal—if accordingly translated to an energy scale—and nominal Nd3+ concentration x is present over the whole regime between x = 0 and x = 1, in agreement with expectations from electromagnetic dispersion theory (see Figure 3b) [67]. On a statistical significance level of α = 0.05, the intercept of the linear calibration line does not differ from zero, as verified by a conventional t test. Altogether, the diffuse reflectance spectra confirm a homogeneous miscibility between the LaPO4 and NdPO4 phases within the nanocrystals.
In the context of suitability for in vivo nanothermometry, the Kubelka-Munk spectra (see Figure 3a) also reveal that the strongest absorption transitions of Nd3+ within the discussed nanocrystals are located in the first biological window between 650 nm and 950 nm. This demonstrates the general suitability of Nd3+ as a potent absorber for in vivo applications since its absorption remains negligibly affected by light attenuation due to surrounding tissue. The linear correlation of the integrated absorption strength with concentration in the La1−xNdxPO4 nanocrystals does in principle allow for a simple strategy to improve the absorption strength in the Nd3+-activated nanocrystals using higher Nd3+ contents. As will be discussed below, however, this can induce both energy migration and cross-relaxation effects that limit the temperature window and thermometric performance of the nanocrystals.

3.3. Photoluminescence Properties and Luminescence Decay Dynamics—Predictions on Consequences for Thermometry with Nd3+

Figure 4a depicts the photoluminescence emission spectra of the La1−xNdxPO4 nanocrystals at room temperature upon CW laser excitation at 690 nm into the 4F9/2 spin-orbit level of the Nd3+ ions. Irrespective of the Nd3+ content, the room temperature emission spectra show radiative transitions from the 4F3/2 level into the ground levels 4I9/2 ( λ em = 890 nm), 4I11/2 ( λ em = 1063 nm) and 4I13/2 ( λ em = 1341 nm). The 4F3/24I11/2 transition has the largest intensity, reflecting a large branching ratio of around β11/2(4F3/2) = 0.67 compared to the other observable radiative transitions (β9/2(4F3/2) = 0.17, β13/2(4F3/2) = 0.16). This is understandable as | Δ L | = 3, | Δ J | = 4 implies a Judd-Ofelt allowed forced electric dipole transition and moreover, Nd3+ is typically characterized by large Ω4 and Ω6 Judd-Ofelt intensity parameters in phosphates [68,69]. Similar findings have been reported in the previously mentioned higher condensed phosphates ALa1-yNdy(PO3)4 (A = Li – Rb; 0 ≤ y ≤ 1) [66]. As expected, the emission spectra become inhomogeneously broadened and the spectral resolution of the different crystal field components decreases with increasing Nd3+ content. The corresponding photoluminescence excitation spectra at room temperature acquired upon monitoring the most intense 4F3/24I11/2 emissive transition of Nd3+ at 1057 nm are depicted in Figure 4b). While the Kubelka-Munk spectra clearly suggest that the 4I9/24F5/2 transition at λ abs = 808 nm shows the maximum absorption coefficient, as expected for a Judd-Ofelt-allowed transition ( | Δ L | = 3, | Δ J | = 2), the excitation spectra reveal that absorption into any of the 4FJ (J = 3/2…9/2) levels efficiently induces emission from the 4F3/2 level at room temperature, in particular with increasing Nd3+ content. Like in the emission spectra, the respective excitation transitions also suffer inhomogeneous broadening. Moreover, several artefacts due to the Xe lamp lines are present in the excitation spectra (see peaks marked with asterisks in Figure 4b). These artefacts can serve as an internal intensity standard. The strong decrease in relative intensity of Nd3+ excitation lines relative to the Xe-lamp lines with x indicates the presence of concentration quenching of the photoluminescence in the nanocrystals.
Important information on radiative and non-radiative decay rates can be obtained from luminescence decay measurements. Figure 5a depicts the luminescence decay curves obtained upon direct excitation into the 4F3/2 level of Nd3+ at 870 nm and detection of the most intense 4F3/24I11/2 emissive transition around 1060 nm. The intensity of the 4F3/2–excited emission at λem = 1058 nm in a very dilute, microcrystalline La0.999Nd0.001PO4 control sample (λex = 870 nm) decays monoexponentially with an (assumed purely radiative) decay rate of kr(4F3/2) = 2.25 ms−1 at room temperature (see Figure 5a). With increasing Nd3+ content, a faster and non-exponential decay is observed. This behavior is a clear signature of an energy transfer processes involving quenching of the 4F3/2 emission. In order to gain insight in the underlying quenching mechanism, the average decay rates k were determined with Equation (2) from the decay data depicted in Figure 5a.
k = 1 τ = j I ( t j ) j t j   ×   I ( t j )
In Equation (2), I(tj) denotes the normalized, background-corrected luminescence intensity at time tj and j runs over all acquired data points. In the case of NdPO4, the luminescence decay is already so fast that it was not feasible to reliably determine an average decay rate (see Figure 5a). Thus, it was excluded from further analysis. A plot of the average decay rates in La1−xNdxPO4 versus the Nd3+ concentration x reveals an approximately linear relation between the two quantities (see Figure 5b). This is a clear signature of a two-ion process [70] and indicates that cross-relaxation of the 4F3/2 level between the Nd3+ ions becomes active in the LaPO4 host even at concentrations as low as 2 mol%. An explanation for the high cross-relaxation efficiency is the small nearest neighbor distance of Nd3+ ions in the orthophosphates (only 4.036 Å (NdPO4), 4.104 Å (LaPO4)) [60,61], which allows for an efficient electric dipole—electric dipole-type energy transfer. At lower concentrations and homogeneous doping, the average distances between Nd3+ ions will be larger and thus, the cross-relaxation efficiency is reduced. However, for electric dipole-electric dipole type energy transfer, cross-relaxation can be still effective over distances as large as 10 Å.
Inspection of the energy level diagram of Nd3+ shows that indeed the phonon-assisted cross-relaxation pathways [Nd1, Nd2]: [4F3/2, 4I9/2] → [4I15/2, 4I13/2] + ω eff and [4F3/2, 4I9/2] → [4I13/2, 4I15/2] + ω eff can take place if ω eff 1050–1100 cm−1 (see Figure 5c). The required phonon energy agrees very well with the asymmetric O-P-O stretching vibration [59,66], whose energy also matches the energy gap between the 4F5/2 and 4F3/2 level. The 4F5/2 level is even prone to resonant cross-relaxation, [Nd1, Nd2]: [4F5/2, 4I9/2] → [4I15/2, 4I15/2] (see Figure 5c). The possibility of both the 4F3/2 and 4F5/2 level to decay via cross-relaxation effectively increases the decay rate of both spin-orbit levels. This is expected to have immediate consequences for the thermometric performance of Nd3+ once the average decay rates of those levels become similar to the non-radiative transition rates governing the thermal coupling of the excited states. If the average decay rates supersede the non-radiative transition rates to bridge the 4F5/24F3/2 gap, thermodynamic Boltzmann equilibrium cannot be sustained anymore, and the thermometric performance is lost.
In order to obtain a semi-quantitative measure for the non-radiative transition rates mediating the thermal coupling between the 4F5/2 and 4F3/2 levels of Nd3+, luminescence decay curves upon selective excitation into each of those levels were recorded for the dilute microcrystalline La0.999Nd0.001PO4 sample. For excitation into either the 4F3/2 (λex = 870 nm) or 4F5/2 level (λex = 790 nm) and monitoring the luminescence decay of the 4F3/24I11/2-related transition at 1058 nm at room temperature, a purely single exponential decay with a radiative decay rate of kr(4F3/2) = 2.25 ms−1 is observed with errors well below 0.01 ms−1 (see Figure 6a,b).
This observation indicates that the sum of radiative and non-radiative decay rate from the 4F5/2 level, kr(4F5/2) + k nr em ( T ) , have to be much higher than the radiative decay rate kr(4F3/2) since otherwise a rise component is expected in the decay curve recorded for 4F5/2 excitation. For selective excitation at 870 nm (resonant with excitation into the 4F3/2 level) very weak anti-Stokes 4F5/2 emission is observed with a much faster initial decay component besides the slower decay of the overlapping 4F3/24I9/2 – related luminescence. This fast ~4 μs initial decay is assigned to the total decay rate of the 4F5/2 level. In order to obtain a more precise value, upon selective excitation into the 4F5/2 level, the luminescence decay was recorded for the strongest 4F5/24I9/2 – related emission at 804 nm (see Figure 7a). A fast decay with an average (radiative and non-radiative) decay rate of kr(4F5/2) + k nr em ( T = 298   K ) = 204 ms−1 was found using Equation (2). The value agrees well with the result for the fast component from the double exponential fit (kr(4F5/2) + k nr em ( 298   K ) = (242 ± 12) ms−1) as depicted in Figure 6c. For the following analysis, we will employ the average value kr(4F5/2) + k nr em ( 298   K ) = (223 ± 19) ms−1.
In order to separate the contribution of radiative decay rate kr(4F5/2) from the non-radiative rate k nr em ( T ) at temperature T, an emission spectrum containing the emissive transitions from both the 4F5/2 and 4F3/2 spin-orbit levels into the same ground level 4I13/2 was recorded for selective excitation into the 4F5/2 level (λex = 808 nm). The ratio between the integrated intensities (if measured in photon counts) of the emission, Iem, from the 4F5/2 and the 4F3/2 is equal to the ratio between kr(4F5/2) and k nr em ( T ) , respectively:
k r ( F 4 5 / 2 ) k nr em ( T )   = I em ( F 4 5 / 2 ) I em ( F 4 3 / 2 ) = β ( F 4 3 / 2 I 4 13 / 2 ) I em ( F 4 5 / 2 I 4 13 / 2 ) β ( F 4 5 / 2 I 4 13 / 2 ) I em ( F 4 3 / 2 I 4 13 / 2 )
with β ( F 4 3 / 2 I 4 13 / 2 ) = 0.16 and β ( F 4 5 / 2 I 4 13 / 2 ) = 0.60 as the branching ratios of the respective transitions derived from the luminescence spectra (see Figure 4a and Figure 7b). Equation (3) is only valid if thermally excited non-radiative absorption from the 4F3/2 level back to the 4F5/2 is negligible and if there are no other non-radiative decay paths for the two levels.
The temperature dependence of non-radiative rates among 4fn-related spin-orbit levels is governed by thermally excited multi-phonon transitions. The thermal average number n eff of effective phonon modes with energy ω eff resonantly bridging the regarded energy gap between two electronic levels is given by the Planck formula:
n eff = 1 exp ( ω eff k B T ) 1
with kB as the Boltzmann constant. With an effective phonon energy of ω eff 1050 cm−1, even at room temperature (T = 298 K), it is n eff 6.34 ∙ 10−3, i.e., effectively, one phonon mode is only thermally excited with a very low probability. Based on Equation (3) and the emission spectrum depicted in Figure 7b, a non-radiative emission rate of k nr em (298 K) = g1knr(0) ( 1 + n eff ) g1knr(0) = (219 ± 19) ms−1 can be derived, where g1 = 4 is the (2J + 1)-fold degeneracy of the 4F3/2 level. Thus, the intrinsic non-radiative rate is estimated to be knr(0) = (54.6 ± 4.7) ms−1.
The validity of the employed approximation of a still negligible thermal multi-phonon absorption rate for the non-radiative transition 4F3/24F5/2 can now be verified. With the value for knr(0) and the degeneracy of the 4F5/2 level, g2 = 6, it is k nr abs (298 K) = g2knr(0) n eff = (2.08 ± 0.18) ms−1, which agrees with kr(4F3/2) within the statistical error. Thus, the approximation to neglect that non-radiative absorption pathway for Nd3+ ions in the dilute La0.999Nd0.001PO4 microcrystals is just about to fail at room temperature and Boltzmann behavior should be expected to set in just above room temperature. In the case of the concentrated Nd3+-doped nanocrystals La1−xNdxPO4, however, the additional cross-relaxation pathways and possible quenching of the 4F3/2 and 4F5/2 emission by high energy vibrations of any capping ligands such as -OH groups on the nanocrystal surface effectively increases the decay rate of the 4F3/2 and 4F5/2 level (see Figure 5a). Consequently, Boltzmann equilibrium is then expected to become active only at successively higher temperatures, once k nr abs ( T ) supersedes these higher average decay rates. Table 2 compiles the relevant decay rates derived from decay dynamics and emission spectra of the dilute microcrystalline La0.999Nd0.001PO4 sample.

3.4. Consequences of Cross-Relaxation on Luminescence Thermometry Employing the 4F5/2 and 4F3/2 Spin-Orbit Levels of Nd3+

For Boltzmann-based luminescence thermometry, it is beneficial to select a different auxiliary excited state that feeds the thermally coupled states giving rise to the temperature-dependent luminescence phenomena. Direct excitation into one of those levels provides an additional non-equilibrium component to the excited state population and although steady-state conditions might be retained, the single ion luminescence thermometer is driven out of thermodynamic equilibrium. Thus, despite the high absorption strength of the 4I9/24F5/2 transition at 808 nm, we decided to selectively excite into the 4F9/2 spin-orbit level with a corresponding absorption wavelength of 690 nm for all luminescence thermometry experiments. The temperature-dependent luminescence spectra of the microcrystalline La0.999Nd0.001PO4 upon excitation at 690 nm are depicted in Figure 8.
All samples were investigated over a wide temperature range, also above physiological temperatures, for the sake of better insight into the thermal coupling between the 4F3/2 and 4F5/2 levels and to investigate the predicted thermometric performance at higher temperatures. In the case of thermodynamic equilibrium conditions between the two excited levels, the LIR, R(T), should obey Boltzmann’s law:
R ( T ) = I 20 I 10 = C g 2 g 1 exp ( Δ E 21 k B T ) ,
where we denote the excited 4F3/2 level as state | 1 and the 4F5/2 level as state | 2 in the following. The ground level 4I9/2 is referred to as state | 0 . Then, g1 = 4 and g2 = 6 represent the (2J + 1)-fold degeneracies of the two spin-orbit levels, while ΔE21 = 1020 cm−1 as derived from the Kubelka-Munk spectra is the energy gap between the two excited levels. This value is in good agreement with expectations according to the Dieke—Carnall diagram [71]. kB is Boltzmann’s constant and C is a pre-exponential factor that basically relates the electronic line strengths of the regarded radiative transitions of interest. It can also be estimated from Judd-Ofelt theory [14,69], but will be considered as a fitting parameter in the present study. Figure 9 depicts the thermometric Boltzmann plots derived from the temperature-dependent luminescence spectra of the La1−xNdxPO4 nanocrystals in the wavelength range between 775 nm and 950 nm (i.e., in BW I). All spectra were corrected for constant instrumental background of the NIR PMT and additional blackbody radiation background, B(λ, T), at a given temperature T to obtain meaningful LIR data,
B ( λ ,   T ) = A + 2 D h c 2 λ 5 [ exp ( h c λ k B T )     1 ]
with A and D as free parameters, λ as the wavelength, h as Planck’s constant and c as the light velocity. Due to concentration quenching the luminescence in NdPO4 was so weak, especially at temperatures above 100 °C, that no meaningful information could be extracted and thus, it was excluded from further analysis. The predictions according to the luminescence decay kinetics of Nd3+ can be directly compared to the experimental data. The LIR of the 4FJ4I9/2 (J = 3/2, 5/2) transitions in La0.999Nd0.001PO4 shows Boltzmann behavior over the full temperature range from 30 °C to 500 °C. This observation is consistent with the observation that the non-radiative absorption rate k nr abs (T) governing the non-radiative 4F3/24F5/2 transition is similar to the radiative decay rate of the 4F3/2 level at room temperature and increases at higher temperatures. This will result in Boltzmann equilibrium starting just above room temperature. Overall, thermodynamic equilibrium between the two spin-orbit levels can be sustained over the full temperature range above ~300 K.
In the La1−xNdxPO4 (x > 0.001) nanocrystals, the average decay rate of the 4F3/2 level increases with Nd3+ content due to the additional cross-relaxation pathways between neighboring Nd3+ ions. This additional decay channel competes with the non-radiative absorption rate and can hamper Boltzmann equilibration. Given the known temperature dependence of k nr abs (T) (scaling with the Planck factor in Equation (4)) and the faster decay of the 4F3/2 level (from the decay curves of the 4F3/2 emission in Figure 5), it is possible to determine the threshold temperature Ton, above which Boltzmann behavior is expected in the more concentrated Nd3+-samples. Ton is taken as the temperature at which the non-radiative absorption k nr abs becomes faster than the total decay rate of the 4F3/2 level. In very good agreement with expectations, a shifted onset of the Boltzmann behavior is observed in all higher concentrated La1−xNdxPO4 nanocrystals, both experimentally and theoretically. The predicted onset temperatures Ton derived from the requirement of equal non-radiative absorption and average decay rates from the 4F3/2 level increases with Nd3+ content x and are indicated in Figure 9. For x = 0.02, Boltzmann equilibrium becomes problematic in the physiological temperature window and for x = 0.05, temperature sensing becomes possible only above 450 K. For the higher Nd3+ concentrations (x = 0.25) no Boltzmann behavior is observed even up to 500 °C. The predicted temperature at which the non-radiative absorption rate dominates is even higher, above 900 °C. Thus, although a higher Nd3+ content may increase the absorption efficiency of the nanocrystals (cf. Figure 3b), the efficient cross-relaxation of Nd3+ ions prevents sustainment of a Boltzmann equilibrium for the excited 4F3/2 and 4F5/2 levels for a larger temperature range. As a result, higher Nd3+ contents destroy the promising potential of the large 4F5/24F3/2 gap for Boltzmann thermometry with high sensitivity at physiological temperatures. In contrast, higher temperature thermometry (100 °C–500 °C) is still feasible even for Nd3+ concentrations as high as x = 0.10. Measurements of higher temperatures in BW I (and consequently, also in BW II) with those nanocrystals is however cumbersome since the blackbody background starts to dominate the emission spectrum and temperature can be measured more accurately from the background itself.
The effect of cross-relaxation on the useable temperature window for LIR temperature sensing can differ among lanthanide ions. Here we show for the 4F3/2 and 4F5/2 levels of Nd3+ that high dopant concentrations are clearly detrimental because cross-relaxation shortens the lifetime of the emitting levels, thus limiting the time available for Boltzmann equilibration. However, cross-relaxation can also be beneficial, if it provides an additional pathway for thermalization between the emitting levels, thus establishing Boltzmann equilibrium. This has been shown to be the case for the 5D0 and 5D1 levels of Eu3+ where cross-relaxation between neighboring Eu3+ ions provides an alternative path for relaxation and thus sustains Boltzmann behavior over a wider temperature range at elevated Eu3+ concentrations, as was demonstrated in the case of β-NaYF4:Eu3+ [58].
It is noteworthy that the fitted effective energy gaps, ΔE21, gradually increase with higher Nd3+ content x. This can be explained by the fact that the fitted effective energy gap for the nanocrystals with higher Nd3+ contents are obtained from higher temperature data. At higher temperatures, the probability for non-radiative absorption into the higher crystal field states of the 4F5/2 spin-orbit level increases, which effectively increases the energy gap ΔE21. Thus, only in the very dilute La0.999Nd0.001PO4 compound does the energy gap agree with the spectroscopically deduced energy gap according to the Kubelka-Munk spectra (ΔE21 = 1020 cm−1, see Figure 3a). Interestingly, also the pre-exponential constant C systematically increases with increasing Nd3+ content x. As it is fundamentally related to both the branching ratios and the radiative decay rates from the two considered excited states of Nd3+, its increase may also be related to the temperature-induced population of higher energetic crystal field states of the 4F3/2 and 4F5/2 levels. In conjunction with those observations, the predicted onset temperatures for Boltzmann behavior for the nanocrystals activated with 5 mol% and 10 mol% Nd3+ are significantly higher than the experimentally observed onsets (see Figure 9c,d). A possible explanation is a decreasing cross-relaxation efficiency of the 4F3/2 level at higher temperatures. Since cross-relaxation of the 4F3/2 level requires one high energy phonon mode of the phosphate host, an increasing temperature and population of the higher energy crystal field level may lead to an energy mismatch of the necessary energy transfer resonance condition and could thus reduce the cross-relaxation efficiency. Temperature-dependent luminescence decay analyses and modelling of the energy migration processes are necessary to confirm this hypothesis.
Since thermodynamic equilibrium between the 4F5/2 and 4F3/2 level of Nd3+ is sustained over the full temperature range investigated (30 °C–500 °C) in La0.999Nd0.001PO4, physiological temperatures are measurable by means of luminescence thermometry with that compound. As Boltzmann behavior is realized, the relative sensitivity Sr (in % K−1) of the luminescence thermometer is given by
S r = | 1 R ( T ) d R d T | = Δ E 21 k B T 2
Figure 10 depicts the evolution of the relative sensitivity for La0.999Nd0.001PO4 as obtained from Equation (7). In particular, it is higher than 1% K−1 for the full physiological temperature regime (30 °C–75 °C), which is practically difficult to achieve around room temperature with any single ion Boltzmann thermometer, especially in the NIR regime [49,50,51,57]. Typically, energy transfer-based thermometers are used in those cases [30,43,54,55] for which the underlying thermometric mechanisms are often not well established. The present results show promising potential of Nd3+ for physiological temperature sensing by means of luminescence thermometry, if the boundary conditions for the validity of a Boltzmann equilibrium are met which requires low Nd3+ concentrations. The 4F5/24F3/2 spin-orbit gap in Nd3+ gives rise to relative sensitivities that are an order of magnitude higher than the typically found ones in the range of 0.25% K−1 [50]. A disadvantage of the presented thermometric concept of Nd3+ is the rather low intensity of the 4F5/24I9/2 emission that can give rise to a higher relative intensity uncertainty, depending on the sensitivity of the detection system. Since it is the temperature uncertainty that matters in a well performing luminescence thermometer, both the relative sensitivity and emission intensities have to be optimized.

4. Conclusions

Nd3+ is a promising candidate for in vivo luminescence thermometry due to its intense radiative transitions in the biological transparency windows BW I and BW II. Most Nd-based single ion thermometers utilize the two crystal field states arising from the excited 4F3/2 spin-orbit level. These thermometers are fundamentally limited in their relative thermal sensitivity Sr due to the small energy gap of ~100 cm−1 and the high demands on spectral resolution which are difficult to meet under in vivo conditions. An alternative promising probe for physiological temperature sensing is the 4F5/24F3/2 spin-orbit gap of Nd3+ which is ~1000 cm−1. In order to investigate the feasibility of temperature sensing based on the temperature dependent emission intensity ratio of 4F3/2- and 4F5/2-related emission, both microcrystalline, dilute La0.999Nd0.001PO4 and nanocrystalline La1−xNdxPO4 (x = 0.02, 0.05, 0.10, 0.25, 1.00) were synthesized by means of a co-precipitation approach and structurally, morphologically and optically characterized. Analysis of the decay kinetics reveals that the 4F3/2 and 4F5/2 excited states of Nd3+ are prone to efficient cross-relaxation quenching at higher Nd-concentrations. Cross-relaxation competes with the non-radiative transition rates governing the thermalization of the 4F5/2 and 4F3/2 spin-orbit levels. This additional decay pathway leads to a gradual breakdown of Boltzmann equilibrium at physiological temperatures with increasing Nd3+ concentration. Only for low Nd3+ concentrations (x < 0.02) is it possible to sustain Boltzmann equilibrium between the 4F5/2 and 4F3/2 levels of Nd3+ in the physiological temperature regime. The relative sensitivity in this important temperature regime exceeds 1% K−1. Overall, the present study demonstrates that a careful analysis of the excited state dynamics allows for an assessment of the performance of a luminescence thermometer, but also demonstrates the potential pitfalls of Boltzmann thermometry that can be encountered. A mechanistic understanding of the competition between radiative and non-radiative decay processes can lead to a tailored design of novel luminescence thermometers with optimal performance in a specific temperature window.

Author Contributions

Conceptualization, M.S., Ž.A., V.Ð., S.K., M.D.D., A.M.; data curation, M.S.; formal analysis, M.S.; Ž.A., V.Ð., S.K.; investigation, M.S., Ž.A., V.Ð., S.K.; resources, M.D.D., A.M.; writing—original draft preparation, M.S.; writing—review and editing, M.S., Ž.A., V.Ð., S.K., M.D.D., A.M.; visualization, M.S.; supervision, M.D.D., A.M.; project administration, M.D.D., A.M.; funding acquisition, M.D.D., A.M. All authors have read and agreed to the published version of the manuscript.

Funding

The authors acknowledge funding from the European Union’s Horizon 2020 FET-Open project NanoTBtech (grant agreement No.: 801305).

Acknowledgments

The authors are grateful to J. de Wit and J. D. Meeldijk for assistance during the TEM measurements.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Allison, S.W.; Gillies, G.T. Remote thermometry with thermographic phosphors: Instrumentation and applications. Rev. Sci. Instrum. 1997, 68, 2615–2650. [Google Scholar] [CrossRef]
  2. Cates, M.R.; Beshears, D.L.; Allison, S.W.; Simmons, C.M. Phosphor thermometry at cryogenic temperatures. Rev. Sci. Instrum. 1997, 68, 2412–2417. [Google Scholar] [CrossRef]
  3. Collins, S.F.; Baxter, G.W.; Wade, S.A.; Sun, T.; Grattan, K.T.V.; Zhang, Z.Y.; Palmer, A.W. Comparison of fluorescence-based temperature sensor schemes: Theoretical analysis and experimental validation. J. Appl. Phys. 1998, 84, 4649–4654. [Google Scholar] [CrossRef]
  4. Wade, S.A.; Collins, S.F.; Baxter, G.W. Fluorescence intensity ratio technique for optical fiber point temperature sensing. J. Appl. Phys. 2003, 94, 4743–4756. [Google Scholar] [CrossRef]
  5. Geitenbeek, R.G.; Nieuwelink, A.-E.; Jacobs, T.S.; Salzmann, B.B.V.; Goetze, J.; Meijerink, A.; Weckhuysen, B.M. In Situ Luminescence Thermometry to Locally Measure Temperature Gradients during Catalytic Reactions. ACS Catal. 2018, 8, 2397–2401. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  6. Geitenbeek, R.G.; Vollenbroek, J.C.; Weijgertze, H.M.H.; Tregouet, C.B.M.; Nieuwelink, A.-E.; Kennedy, C.L.; Weckhuysen, B.M.; Lohse, D.; van Blaaderen, A.; van den Berg, A.; et al. Luminescence thermometry for in situ temperature measurements in microfluidic devices. Lab Chip 2019, 19, 1236–1246. [Google Scholar] [CrossRef] [Green Version]
  7. Ravenhorst, I.K.; Geitenbeek, R.G.; Eerden, M.J.; van Tijn Omme, J.; Peréz Garza, H.H.; Meirer, F.; Meijerink, A.; Weckhuysen, B.M. In Situ Local Temperature Mapping in Microscopy Nano-Reactors with Luminescence Thermometry. ChemCatChem 2019, 11, 5505–5512. [Google Scholar] [CrossRef] [Green Version]
  8. Jaque, D.; Vetrone, F. Luminescence nanothermometry. Nanoscale 2012, 4, 4301–4326. [Google Scholar] [CrossRef]
  9. Brites, C.D.S.; Lima, P.P.; Silva, N.J.O.; Millán, A.; Amaral, V.S.; Palacio, F.; Carlos, L.D. Thermometry at the nanoscale. Nanoscale 2012, 4, 4799–4829. [Google Scholar] [CrossRef] [Green Version]
  10. McLaurin, E.J.; Bradshaw, L.R.; Gamelin, D.R. Dual-Emitting Nanoscale Temperature Sensors. Chem. Mater. 2013, 25, 1283–1292. [Google Scholar] [CrossRef]
  11. Wang, X.-D.; Wolfbeis, O.S.; Meier, R.J. Luminescent probes and sensors for temperature. Chem. Soc. Rev. 2013, 42, 7834–7869. [Google Scholar] [CrossRef] [PubMed]
  12. Liu, T.-M.; Conde, J.; Lipiński, T.; Bednarkiewicz, A.; Huang, C.-C. Revisiting the classification of NIR-absorbing/emitting nanomaterials for in vivo bioapplications. NPG Asia Mater. 2016, 8, e295. [Google Scholar] [CrossRef]
  13. del Rosal, B.; Ximendes, E.; Rocha, U.; Jaque, D. In Vivo Luminescence Nanothermometry: From Materials to Applications. Adv. Opt. Mater. 2017, 5, 1600508. [Google Scholar] [CrossRef]
  14. Brites, C.D.S.; Balabhadra, S.; Carlos, L.D. Lanthanide-Based Thermometers: At the Cutting-Edge of Luminescence Thermometry. Adv. Opt. Mater. 2019, 7, 1801239. [Google Scholar] [CrossRef] [Green Version]
  15. Peng, H.; Stich, M.I.J.; Yu, J.; Sun, L.-N.; Fischer, L.H.; Wolfbeis, O.S. Luminescent Europium(III) nanoparticles for sensing and imaging of temperature in the physiological range. Adv. Mater. 2010, 22, 716–719. [Google Scholar] [CrossRef] [PubMed]
  16. Chambers, M.D.; Clarke, D.R. Doped Oxides for High-Temperature Luminescence and Lifetime Thermometry. Annu. Rev. Mater. Res. 2009, 39, 325–359. [Google Scholar] [CrossRef] [Green Version]
  17. Haro-González, P.; Martínez-Maestro, L.; Martín, I.R.; García-Solé, J.; Jaque, D. High-sensitivity fluorescence lifetime thermal sensing based on CdTe quantum dots. Small 2012, 8, 2652–2658. [Google Scholar] [CrossRef]
  18. 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]
  19. Peng, D.; Liu, Y.; Zhao, X.; Kim, K.C. Comparison of lifetime-based methods for 2D phosphor thermometry in high-temperature environment. Meas. Sci. Technol. 2016, 27, 95201. [Google Scholar] [CrossRef]
  20. Dramićanin, M.D.; Milićević, B.; Đorđević, V.; Ristić, Z.; Zhou, J.; Milivojević, D.; Papan, J.; Brik, M.G.; Ma, C.-G.; Srivastava, A.M.; et al. Li2TiO3:Mn4+ Deep-Red Phosphor for the Lifetime-Based Luminescence Thermometry. ChemistrySelect 2019, 4, 7067–7075. [Google Scholar] [CrossRef]
  21. Mendieta, A.; Fond, B.; Dragomirov, P.; Beyrau, F. A delayed gating approach for interference-free ratio-based phosphor thermometry. Meas. Sci. Technol. 2019, 30, 74002. [Google Scholar] [CrossRef]
  22. Souza, A.S.; Nunes, L.A.O.; Silva, I.G.N.; Oliveira, F.A.M.; da Luz, L.L.; Brito, H.F.; Felinto, M.C.F.C.; Ferreira, R.A.S.; Júnior, S.A.; Carlos, L.D.; et al. Highly-sensitive Eu3+ ratiometric thermometers based on excited state absorption with predictable calibration. Nanoscale 2016, 8, 5327–5333. [Google Scholar] [CrossRef] [PubMed]
  23. Marciniak, L.; Bednarkiewicz, A.; Elzbieciak, K. NIR–NIR photon avalanche based luminescent thermometry with Nd3+ doped nanoparticles. J. Mater. Chem. C 2018, 6, 7568–7575. [Google Scholar] [CrossRef]
  24. Trejgis, K.; Bednarkiewicz, A.; Marciniak, L. Engineering excited state absorption based nanothermometry for temperature sensing and imaging. Nanoscale 2020. [Google Scholar] [CrossRef] [PubMed]
  25. Rocha, J.; Brites, C.D.S.; Carlos, L.D. Lanthanide Organic Framework Luminescent Thermometers. Chem. Eur. J. 2016, 22, 14782–14795. [Google Scholar] [CrossRef] [PubMed]
  26. Kaczmarek, A.M.; Liu, Y.-Y.; Kaczmarek, M.K.; Liu, H.; Artizzu, F.; Carlos, L.D.; van der Voort, P. Developing Luminescent Ratiometric Thermometers Based on a Covalent Organic Framework (COF). Angew. Chem. 2020, 132, 1948–1956. [Google Scholar] [CrossRef] [Green Version]
  27. Smith, A.M.; Mancini, M.C.; Nie, S. Bioimaging: Second window for in vivo imaging. Nat. Nanotechnol. 2009, 4, 710–711. [Google Scholar] [CrossRef] [Green Version]
  28. Hemmer, E.; Venkatachalam, N.; Hyodo, H.; Hattori, A.; Ebina, Y.; Kishimoto, H.; Soga, K. Upconverting and NIR emitting rare earth based nanostructures for NIR-bioimaging. Nanoscale 2013, 5, 11339–11361. [Google Scholar] [CrossRef]
  29. Hemmer, E.; Benayas, A.; Légaré, F.; Vetrone, F. Exploiting the biological windows: Current perspectives on fluorescent bioprobes emitting above 1000 nm. Nanoscale Horiz. 2016, 1, 168–184. [Google Scholar] [CrossRef]
  30. Ximendes, E.C.; Santos, W.Q.; Rocha, U.; Kagola, U.K.; Sanz-Rodríguez, F.; Fernández, N.; Gouveia-Neto, A.d.S.; Bravo, D.; Domingo, A.M.; del Rosal, B.; et al. Unveiling in Vivo Subcutaneous Thermal Dynamics by Infrared Luminescent Nanothermometers. Nano Lett. 2016, 16, 1695–1703. [Google Scholar] [CrossRef]
  31. Ximendes, E.C.; Rocha, U.; del Rosal, B.; Vaquero, A.; Sanz-Rodríguez, F.; Monge, L.; Ren, F.; Vetrone, F.; Ma, D.; García-Solé, J.; et al. In Vivo Ischemia Detection by Luminescent Nanothermometers. Adv. Healthc. Mater. 2017, 6, 1601195. [Google Scholar] [CrossRef] [PubMed]
  32. Ximendes, E.C.; Rocha, U.; Sales, T.O.; Fernández, N.; Sanz-Rodríguez, F.; Martín, I.R.; Jacinto, C.; Jaque, D. In Vivo Subcutaneous Thermal Video Recording by Supersensitive Infrared Nanothermometers. Adv. Funct. Mater. 2017, 27, 1702249. [Google Scholar] [CrossRef]
  33. Santos, H.D.A.; Ximendes, E.C.; La Iglesias-de Cruz, M.d.C.; Chaves-Coira, I.; del Rosal, B.; Jacinto, C.; Monge, L.; Rubia-Rodríguez, I.; Ortega, D.; Mateos, S.; et al. In Vivo Early Tumor Detection and Diagnosis by Infrared Luminescence Transient Nanothermometry. Adv. Funct. Mater. 2018, 28, 1803924. [Google Scholar] [CrossRef]
  34. del Rosal, B.; Ruiz, D.; Chaves-Coira, I.; Juárez, B.H.; Monge, L.; Hong, G.; Fernández, N.; Jaque, D. In Vivo Contactless Brain Nanothermometry. Adv. Funct. Mater. 2018, 28, 1806088. [Google Scholar] [CrossRef] [Green Version]
  35. Laha, S.S.; Naik, A.R.; Kuhn, E.R.; Alvarez, M.; Sujkowski, A.; Wessells, R.J.; Jena, B.P. Nanothermometry Measure of Muscle Efficiency. Nano Lett. 2017, 17, 1262–1268. [Google Scholar] [CrossRef] [PubMed]
  36. Ortgies, D.H.; García-Villalón, Á.L.; Granado, M.; Amor, S.; Rodríguez, E.M.; Santos, H.D.A.; Yao, J.; Rubio-Retama, J.; Jaque, D. Infrared fluorescence imaging of infarcted hearts with Ag2S nanodots. Nano Res. 2019, 12, 749–757. [Google Scholar] [CrossRef]
  37. Shen, Y.; Lifante, J.; Ximendes, E.; Santos, H.D.A.; Ruiz, D.; Juárez, B.H.; Zabala Gutiérrez, I.; Torres Vera, V.; Rubio Retama, J.; Martín Rodríguez, E.; et al. Perspectives for Ag2S NIR-II nanoparticles in biomedicine: From imaging to multifunctionality. Nanoscale 2019, 11, 19251–19264. [Google Scholar] [CrossRef]
  38. Rocha, U.; Jacinto da Silva, C.; Ferreira Silva, W.; Guedes, I.; Benayas, A.; Martínez Maestro, L.; Acosta Elias, M.; Bovero, E.; van Veggel, F.C.J.M.; García Solé, J.A.; et al. Subtissue thermal sensing based on neodymium-doped LaF3 nanoparticles. ACS Nano 2013, 7, 1188–1199. [Google Scholar] [CrossRef]
  39. Marciniak, L.; Bednarkiewicz, A.; Kowalska, D.; Strek, W. A new generation of highly sensitive luminescent thermometers operating in the optical window of biological tissues. J. Mater. Chem. C 2016, 4, 5559–5563. [Google Scholar] [CrossRef]
  40. Rocha, U.; Jacinto, C.; Kumar, K.U.; López, F.J.; Bravo, D.; Solé, J.G.; Jaque, D. Real-time deep-tissue thermal sensing with sub-degree resolution by thermally improved Nd3+:LaF3 multifunctional nanoparticles. J. Lumin. 2016, 175, 149–157. [Google Scholar] [CrossRef]
  41. Vetrone, F.; Naccache, R.; Zamarrón, A.; La Juarranz de Fuente, A.; Sanz-Rodríguez, F.; Martinez Maestro, L.; Martín Rodriguez, E.; Jaque, D.; García Solé, J.; Capobianco, J.A. Temperature sensing using fluorescent nanothermometers. ACS Nano 2010, 4, 3254–3258. [Google Scholar] [CrossRef] [PubMed]
  42. Skripka, A.; Benayas, A.; Marin, R.; Canton, P.; Hemmer, E.; Vetrone, F. Double rare-earth nanothermometer in aqueous media: Opening the third optical transparency window to temperature sensing. Nanoscale 2017, 9, 3079–3085. [Google Scholar] [CrossRef] [Green Version]
  43. Mi, C.; Zhou, J.; Wang, F.; Lin, G.; Jin, D. Ultra-Sensitive Ratiometric Nanothermometer with Large Dynamic Range and Photostability. Chem. Mater. 2019. [Google Scholar] [CrossRef]
  44. Wawrzynczyk, D.; Bednarkiewicz, A.; Nyk, M.; Strek, W.; Samoc, M. Neodymium(III) doped fluoride nanoparticles as non-contact optical temperature sensors. Nanoscale 2012, 4, 6959–6961. [Google Scholar] [CrossRef]
  45. Marciniak, L.; Prorok, K.; Bednarkiewicz, A.; Kowalczyk, A.; Hreniak, D.; Strek, W. Water dispersible LiNdP4O12 nanocrystals: New multifunctional NIR–NIR luminescent materials for bio-applications. J. Lumin. 2016, 176, 144–148. [Google Scholar] [CrossRef]
  46. Boyer, J.-C.; Vetrone, F.; Cuccia, L.A.; Capobianco, J.A. Synthesis of colloidal upconverting NaYF4 nanocrystals doped with Er3+, Yb3+ and Tm3+, Yb3+ via thermal decomposition of lanthanide trifluoroacetate precursors. J. Am. Chem. Soc. 2006, 128, 7444–7445. [Google Scholar] [CrossRef]
  47. Boyer, J.-C.; Cuccia, L.A.; Capobianco, J.A. Synthesis of Colloidal Upconverting NaYF4:Er3+/Yb3+ and Tm3+/Yb3+ Monodisperse Nanocrystals. Nano Lett. 2007, 7, 847–852. [Google Scholar] [CrossRef]
  48. Wang, F.; Deng, R.; Liu, X. Preparation of core-shell NaGdF4 nanoparticles doped with luminescent lanthanide ions to be used as upconversion-based probes. Nat. Protoc. 2014, 9, 1634–1644. [Google Scholar] [CrossRef]
  49. Benayas, A.; del Rosal, B.; Pérez-Delgado, A.; Santacruz-Gómez, K.; Jaque, D.; Hirata, G.A.; Vetrone, F. Nd:YAG Near-Infrared Luminescent Nanothermometers. Adv. Opt. Mater. 2015, 3, 687–694. [Google Scholar] [CrossRef]
  50. Dantelle, G.; Matulionyte, M.; Testemale, D.; Cantarano, A.; Ibanez, A.; Vetrone, F. Nd3+ doped Gd3Sc2Al3O12 nanoparticles: Towards efficient nanoprobes for temperature sensing. Phys. Chem. Chem. Phys. 2019, 21, 11132–11141. [Google Scholar] [CrossRef]
  51. Skripka, A.; Morinvil, A.; Matulionyte, M.; Cheng, T.; Vetrone, F. Advancing neodymium single-band nanothermometry. Nanoscale 2019, 11, 11322–11330. [Google Scholar] [CrossRef] [PubMed]
  52. Dantelle, G.; Testemale, D.; Homeyer, E.; Cantarano, A.; Kodjikian, S.; Dujardin, C.; Hazemann, J.-L.; Ibanez, A. A new solvothermal method for the synthesis of size-controlled YAG:Ce single-nanocrystals. RSC Adv. 2018, 8, 26857–26870. [Google Scholar] [CrossRef] [Green Version]
  53. Suta, M.; Meijerink, A. A generalized theoretical framework of ratiometric single ion luminescent thermometers: Quantitative guidelines and limitations of the Boltzmann distribution for an optimum choice. 2020. to be submitted. [Google Scholar]
  54. Marciniak, Ł.; Bednarkiewicz, A.; Stefanski, M.; Tomala, R.; Hreniak, D.; Strek, W. Near infrared absorbing near infrared emitting highly-sensitive luminescent nanothermometer based on Nd3+ to Yb3+ energy transfer. Phys. Chem. Chem. Phys. 2015, 17, 24315–24321. [Google Scholar] [CrossRef] [PubMed]
  55. Marciniak, L.; Prorok, K.; Francés-Soriano, L.; Pérez-Prieto, J.; Bednarkiewicz, A. A broadening temperature sensitivity range with a core-shell YbEr@YbNd double ratiometric optical nanothermometer. Nanoscale 2016, 8, 5037–5042. [Google Scholar] [CrossRef] [PubMed]
  56. Haro-González, P.; Martín, I.R.; Martín, L.L.; León-Luis, S.F.; Pérez-Rodríguez, C.; Lavín, V. Characterization of Er3+ and Nd3+ doped Strontium Barium Niobate glass ceramic as temperature sensors. Opt. Mater. 2011, 33, 742–745. [Google Scholar] [CrossRef]
  57. Kolesnikov, I.E.; Kalinichev, A.A.; Kurochkin, M.A.; Mamonova, D.V.; Kolesnikov, E.Y.; Lähderanta, E.; Mikhailov, M.D. Bifunctional heater-thermometer Nd3+-doped nanoparticles with multiple temperature sensing parameters. Nanotechnology 2019, 30, 145501. [Google Scholar] [CrossRef]
  58. Geitenbeek, R.G.; de Wijn, H.W.; Meijerink, A. Non-Boltzmann Luminescence in NaYF4:Eu3+: Implications for Luminescence Thermometry. Phys. Rev. Appl. 2018, 10, 64006. [Google Scholar] [CrossRef] [Green Version]
  59. Silva, E.N.; Ayala, A.P.; Guedes, I.; Paschoal, C.W.A.; Moreira, R.L.; Loong, C.-K.; Boatner, L.A. Vibrational spectra of monazite-type rare-earth orthophosphates. Opt. Mater. 2006, 29, 224–230. [Google Scholar] [CrossRef]
  60. Ni, Y.; Hughes, J.M.; Mariano, A.N. Crystal chemistry of the monazite and xenotime structures. Am. Mineral. 1995, 80, 21–26. [Google Scholar] [CrossRef]
  61. Mullica, D.F.; Grossie, D.A.; Boatner, L.A. Structural refinements of praseodymium and neodymium orthophosphate. J. Solid State Chem. 1985, 58, 71–77. [Google Scholar] [CrossRef]
  62. Shannon, R.D. Revised Effective Ionic Radii and Systematic Studies of Interatomic Distances in Halides and Chalcogenides. Acta Cryst. A 1976, 32, 751–767. [Google Scholar] [CrossRef]
  63. Li, W.; Zhang, Q.; Joos, J.J.; Smet, P.F.; Schmedt auf der Günne, J. Blind spheres of paramagnetic dopants in solid state NMR. Phys. Chem. Chem. Phys. 2019, 21, 10185–10194. [Google Scholar] [CrossRef] [Green Version]
  64. Li, W.; Adlung, M.; Zhang, Q.; Wickleder, C.; Schmedt auf der Günne, J. A Guide to Brighter Phosphors-Linking Luminescence Properties to Doping Homogeneity Probed by NMR. ChemPhysChem 2019, 20, 3245–3250. [Google Scholar] [CrossRef] [Green Version]
  65. Li, W.; Smet, P.F.; Martin, L.I.D.J.; Pritzel, C.; Schmedt auf der Günne, J. Doping homogeneity in co-doped materials investigated at different length scales. Phys. Chem. Chem. Phys. 2020, 22, 818–825. [Google Scholar] [CrossRef] [Green Version]
  66. Marciniak, L.; Strek, W.; Guyot, Y.; Hreniak, D.; Boulon, G. Synthesis and Nd3+ Luminescence Properties of ALa1−xNdxP4O12 (A = Li, Na, K, Rb) Tetraphosphate Nanocrystals. J. Phys. Chem. C 2015, 119, 5160–5167. [Google Scholar] [CrossRef]
  67. Mayerhöfer, T.G.; Pipa, A.V.; Popp, J. Beer’s Law-Why Integrated Absorbance Depends Linearly on Concentration. ChemPhysChem 2019, 20, 2748–2753. [Google Scholar] [CrossRef]
  68. Görller-Walrand, C.; Binnemans, K. Spectral Intensities of f-f Transitions: Chapter 167. In Handbook on the Physics and Chemistry of Rare Earths; Gschneidner, K.A., Jr., Eyring, L., Eds.; Elsevier: Amsterdam, The Netherlands, 1998; pp. 101–264. [Google Scholar] [CrossRef]
  69. Ćirić, A.; Stojadinović, S.; Dramićanin, M.D. An extension of the Judd-Ofelt theory to the field of lanthanide thermometry. J. Lumin. 2019, 216, 116749. [Google Scholar] [CrossRef]
  70. Fong, F.K.; Diestler, D.J. Many-Body Processes in Nonradiative Energy Transfer between Ions in Crystals. J. Chem. Phys. 1972, 56, 2875–2880. [Google Scholar] [CrossRef]
  71. Carnall, W.T.; Goodman, G.L.; Rajnak, K.; Rana, R.S. A systematic analysis of the spectra of the lanthanides doped into single crystal LaF3. J. Chem. Phys. 1989, 90, 3443–3457. [Google Scholar] [CrossRef]
Figure 1. (a) X-ray powder diffraction (XRPD) patterns of synthesized microcrystalline La0.999Nd0.001PO4 and La1−xNdxPO4 nanocrystals in comparison to the diffraction patterns of LaPO4 (ICSD: 79747) and NdPO4 (ICSD: 62162) as simulated from single crystal diffraction data [60,61]. The reflections marked with asterisks are due to the employed Al sample holder. The diffraction patterns were stacked for better comparison. (b) Enlarged view on the XRPD patterns of the La1−xNdxPO4 nanocrystals over a selected 2θ range (24°–33°) to emphasize the gradual shift of reflections towards higher angles with increasing x.
Figure 1. (a) X-ray powder diffraction (XRPD) patterns of synthesized microcrystalline La0.999Nd0.001PO4 and La1−xNdxPO4 nanocrystals in comparison to the diffraction patterns of LaPO4 (ICSD: 79747) and NdPO4 (ICSD: 62162) as simulated from single crystal diffraction data [60,61]. The reflections marked with asterisks are due to the employed Al sample holder. The diffraction patterns were stacked for better comparison. (b) Enlarged view on the XRPD patterns of the La1−xNdxPO4 nanocrystals over a selected 2θ range (24°–33°) to emphasize the gradual shift of reflections towards higher angles with increasing x.
Nanomaterials 10 00543 g001
Figure 2. Transmission electron microscopy (TEM) images of the synthesized La1−xNdxPO4 nanocrystals: (a) x = 0.02, (b) x = 0.05, (c) x = 0.10, (d) x = 0.25 and (e) x = 1.00. Scale bars indicate either 20 nm or 50 nm, respectively. Selected single nanorods are marked by white dashed ellipses.
Figure 2. Transmission electron microscopy (TEM) images of the synthesized La1−xNdxPO4 nanocrystals: (a) x = 0.02, (b) x = 0.05, (c) x = 0.10, (d) x = 0.25 and (e) x = 1.00. Scale bars indicate either 20 nm or 50 nm, respectively. Selected single nanorods are marked by white dashed ellipses.
Nanomaterials 10 00543 g002
Figure 3. (a) Kubelka-Munk graphs of the La1−xNdxPO4 (x = 0.02, 0.05, 0.10, 0.25, 1.00) nanocrystals in powder form as obtained from the respective diffuse reflectance spectra (see Equation (1)). The respective assignment of the absorption transitions and the first biological window (BW I) are indicated. (b) Correlation between the Nd3+ content x and the integrated absorption signal when evaluated in wavenumber scales. The red dashed line indicates a least-squares linear fit to the data. The intercept does not differ from zero at the statistical significance level of α = 0.05.
Figure 3. (a) Kubelka-Munk graphs of the La1−xNdxPO4 (x = 0.02, 0.05, 0.10, 0.25, 1.00) nanocrystals in powder form as obtained from the respective diffuse reflectance spectra (see Equation (1)). The respective assignment of the absorption transitions and the first biological window (BW I) are indicated. (b) Correlation between the Nd3+ content x and the integrated absorption signal when evaluated in wavenumber scales. The red dashed line indicates a least-squares linear fit to the data. The intercept does not differ from zero at the statistical significance level of α = 0.05.
Nanomaterials 10 00543 g003
Figure 4. (a) Photoluminescence spectra of the powdered La1−xNdxPO4 nanocrystals acquired upon laser excitation with 690 nm at room temperature. The respective transitions and biological windows are indicated. The microcrystalline La0.999Nd0.001PO4 control sample was also added. (b) Photoluminescence excitation spectra of the powdered La1−xNdxPO4 nanocrystals acquired upon monitoring the 4F3/24I11/2 radiative transition at 1057 nm at room temperature. The respective transitions and biological window are indicated. The microcrystalline La0.999Nd0.001PO4 control sample was also added. Peaks marked with an asterisk stem from the employed Xe lamp and could not be removed even by usage of a 1020 nm long pass filter. All spectra were stacked for the sake of clarity.
Figure 4. (a) Photoluminescence spectra of the powdered La1−xNdxPO4 nanocrystals acquired upon laser excitation with 690 nm at room temperature. The respective transitions and biological windows are indicated. The microcrystalline La0.999Nd0.001PO4 control sample was also added. (b) Photoluminescence excitation spectra of the powdered La1−xNdxPO4 nanocrystals acquired upon monitoring the 4F3/24I11/2 radiative transition at 1057 nm at room temperature. The respective transitions and biological window are indicated. The microcrystalline La0.999Nd0.001PO4 control sample was also added. Peaks marked with an asterisk stem from the employed Xe lamp and could not be removed even by usage of a 1020 nm long pass filter. All spectra were stacked for the sake of clarity.
Nanomaterials 10 00543 g004
Figure 5. (a) Luminescence decay curves upon monitoring the 4F3/24I11/2 transition of the La1−xNdxPO4 nanocrystals and the microcrystalline La0.999Nd0.001PO4 control sample at room temperature using pulsed direct excitation into the 4F3/2 level. The decay curve of a microcrystalline dilute La0.999Nd0.001PO4 powder is also depicted for comparison. (b) Plot of the average decay rates k obtained for the La1−xNdxPO4 nanocrystals versus Nd3+ content x. The NdPO4 sample is excluded from the analysis. The red dashed line indicates a least-squares linear fit to the data. The intercept does not differ from zero at the statistical significance level of α = 0.05. (c) Schematic overview over the possible (eventually phonon-assisted) cross-relaxation processes that Nd3+ can undergo at higher concentrations. Blue arrows refer to cross-relaxation of the 4F5/2 level, brown and red arrows to cross-relaxation of the 4F3/2 level and curly arrows indicate emission or absorption of a resonant phonon.
Figure 5. (a) Luminescence decay curves upon monitoring the 4F3/24I11/2 transition of the La1−xNdxPO4 nanocrystals and the microcrystalline La0.999Nd0.001PO4 control sample at room temperature using pulsed direct excitation into the 4F3/2 level. The decay curve of a microcrystalline dilute La0.999Nd0.001PO4 powder is also depicted for comparison. (b) Plot of the average decay rates k obtained for the La1−xNdxPO4 nanocrystals versus Nd3+ content x. The NdPO4 sample is excluded from the analysis. The red dashed line indicates a least-squares linear fit to the data. The intercept does not differ from zero at the statistical significance level of α = 0.05. (c) Schematic overview over the possible (eventually phonon-assisted) cross-relaxation processes that Nd3+ can undergo at higher concentrations. Blue arrows refer to cross-relaxation of the 4F5/2 level, brown and red arrows to cross-relaxation of the 4F3/2 level and curly arrows indicate emission or absorption of a resonant phonon.
Nanomaterials 10 00543 g005aNanomaterials 10 00543 g005b
Figure 6. Luminescence decay curves of the Nd3+-related luminescence in microcrystalline La0.999Nd0.001PO4 powder. The respective excitation and emission wavelengths are given. Solid red lines depict least-squares fitting curves. The relevant statistical parameters indicating the quality of the fits are also indicated.
Figure 6. Luminescence decay curves of the Nd3+-related luminescence in microcrystalline La0.999Nd0.001PO4 powder. The respective excitation and emission wavelengths are given. Solid red lines depict least-squares fitting curves. The relevant statistical parameters indicating the quality of the fits are also indicated.
Nanomaterials 10 00543 g006
Figure 7. (a) Detailed view on the fast decay component from the 4F5/2 spin-orbit level of Nd3+ in microcrystalline La0.999Nd0.001PO4 powder. (b) High resolution emission spectrum depicting transitions from both the 4F5/2 and 4F3/2 level into the same 4I13/2 ground level acquired upon selected excitation into the 4F5/2 level (λex = 808 nm) to separate the contribution of the non-radiative emission from the radiative emission on the decay rate of the 4F5/2 level.
Figure 7. (a) Detailed view on the fast decay component from the 4F5/2 spin-orbit level of Nd3+ in microcrystalline La0.999Nd0.001PO4 powder. (b) High resolution emission spectrum depicting transitions from both the 4F5/2 and 4F3/2 level into the same 4I13/2 ground level acquired upon selected excitation into the 4F5/2 level (λex = 808 nm) to separate the contribution of the non-radiative emission from the radiative emission on the decay rate of the 4F5/2 level.
Nanomaterials 10 00543 g007
Figure 8. Exemplary temperature-dependent emission spectrum of La0.999Nd0.001PO4 in the spectral range of the first biological window (BW I). The thermally coupled radiative transitions are indicated.
Figure 8. Exemplary temperature-dependent emission spectrum of La0.999Nd0.001PO4 in the spectral range of the first biological window (BW I). The thermally coupled radiative transitions are indicated.
Nanomaterials 10 00543 g008
Figure 9. Semi-log plots of the temperature-dependent luminescence intensity ratio (LIR) between the 4F5/24I9/2- and 4F3/24I9/2-related emission of Nd3+ in (a) microcrystalline La0.999Nd0.001PO4, (b) nanocrystalline La0.98Nd0.02PO4, (c) La0.95Nd0.05PO4, (d) La0.90Nd0.10PO4 and (e) La0.75Nd0.25PO4. The NdPO4 nanocrystals were excluded from the analysis due to dominance of blackbody background already above 100 °C. Least-squares fits to the Boltzmann calibration law (see Equation (5)) are indicated by red solid lines together with the relevant statistical figures of merit. The predicted onset temperatures for Boltzmann equilibrium based on the decay analysis in Section 3.3 (see Figure 5a and Table 2) are also indicated.
Figure 9. Semi-log plots of the temperature-dependent luminescence intensity ratio (LIR) between the 4F5/24I9/2- and 4F3/24I9/2-related emission of Nd3+ in (a) microcrystalline La0.999Nd0.001PO4, (b) nanocrystalline La0.98Nd0.02PO4, (c) La0.95Nd0.05PO4, (d) La0.90Nd0.10PO4 and (e) La0.75Nd0.25PO4. The NdPO4 nanocrystals were excluded from the analysis due to dominance of blackbody background already above 100 °C. Least-squares fits to the Boltzmann calibration law (see Equation (5)) are indicated by red solid lines together with the relevant statistical figures of merit. The predicted onset temperatures for Boltzmann equilibrium based on the decay analysis in Section 3.3 (see Figure 5a and Table 2) are also indicated.
Nanomaterials 10 00543 g009
Figure 10. Temperature evolution of the relative sensitivity of Nd3+ in La0.999Nd0.001PO4 within the Boltzmann validity regime and upon usage of the 4F5/24F3/2 gap for Boltzmann-based luminescence thermometry. The grey dashed line indicates the desirable minimum relative sensitivity for in vivo luminescence thermometry, while the blue dashed regime marks the reported maximum achievable relative sensitivity for thermal coupling between the Kramers’ doublets of the 4F3/2 spin-orbit level [49,50,51].
Figure 10. Temperature evolution of the relative sensitivity of Nd3+ in La0.999Nd0.001PO4 within the Boltzmann validity regime and upon usage of the 4F5/24F3/2 gap for Boltzmann-based luminescence thermometry. The grey dashed line indicates the desirable minimum relative sensitivity for in vivo luminescence thermometry, while the blue dashed regime marks the reported maximum achievable relative sensitivity for thermal coupling between the Kramers’ doublets of the 4F3/2 spin-orbit level [49,50,51].
Nanomaterials 10 00543 g010
Table 1. Average crystallite size estimates d , lattice parameters a, b, c, cell volume V and strain of nanocrystals according to Rietveld refinement on the XRPD patterns depicted in Figure 1. The numbers in brackets refer to the errors in the last digits, respectively. The quality parameters for the Rietveld refinement are given below: Profile factor Rp, weighted profile factor Rwp and optimum expected Rexp factor together with the final goodness of the fit (G.o.f.) to the structural input of the monazite structure type.
Table 1. Average crystallite size estimates d , lattice parameters a, b, c, cell volume V and strain of nanocrystals according to Rietveld refinement on the XRPD patterns depicted in Figure 1. The numbers in brackets refer to the errors in the last digits, respectively. The quality parameters for the Rietveld refinement are given below: Profile factor Rp, weighted profile factor Rwp and optimum expected Rexp factor together with the final goodness of the fit (G.o.f.) to the structural input of the monazite structure type.
La0.98Nd0.02PO4La0.95Nd0.05PO4La0.90Nd0.10PO4La0.75Nd0.25PO4NdPO4
d /nm9.239(17)6.852(18)6.652(6)9.057(19)6.199(18)
a6.8631(13)6.8504(14)6.8592(19)6.8390(15)6.8380(4)
b7.1043(13)7.0901(13)7.0924(19)7.0799(15)6.9890(4)
c6.5290(12)6.5171(12)6.5141(18)6.5053(14)6.4210(4)
V3309.84(21)308.08(39)308.44(23)306.57(26)298.67(11)
Strain %0.37(2)0.39(5)0.34(2)0.50(8)0.42(6)
Rp/%5.796.226.585.937.80
Rwp/%4.364.614.854.475.78
Rexp/%3.843.943.813.683.13
G.o.f.1.501.581.721.612.49
Table 2. Derived radiative and non-radiative rates characterizing the thermal coupling of the 4F3/2 and 4F5/2 spin-orbit levels of Nd3+ in microcrystalline La0.999Nd0.001PO4.
Table 2. Derived radiative and non-radiative rates characterizing the thermal coupling of the 4F3/2 and 4F5/2 spin-orbit levels of Nd3+ in microcrystalline La0.999Nd0.001PO4.
kr(4F3/2)/ms1kr(4F5/2)/ms−1knr(0)/ms−1 k nr em (298 K)/ms−1 k nr abs (298 K)/ms−1
2.253.16 ± 0.2754.6 ± 4.7219 ± 192.08 ± 0.18

Share and Cite

MDPI and ACS Style

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. Nanomaterials 2020, 10, 543. https://0-doi-org.brum.beds.ac.uk/10.3390/nano10030543

AMA Style

Suta M, Antić Ž, Ðorđević V, Kuzman S, Dramićanin MD, Meijerink A. Making Nd3+ a Sensitive Luminescent Thermometer for Physiological Temperatures—An Account of Pitfalls in Boltzmann Thermometry. Nanomaterials. 2020; 10(3):543. https://0-doi-org.brum.beds.ac.uk/10.3390/nano10030543

Chicago/Turabian Style

Suta, Markus, Željka Antić, Vesna Ðorđević, Sanja Kuzman, Miroslav D. Dramićanin, and Andries Meijerink. 2020. "Making Nd3+ a Sensitive Luminescent Thermometer for Physiological Temperatures—An Account of Pitfalls in Boltzmann Thermometry" Nanomaterials 10, no. 3: 543. https://0-doi-org.brum.beds.ac.uk/10.3390/nano10030543

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