Expand this Topic clickable element to expand a topic
Skip to content
Optica Publishing Group

Laser heating of the Y1-xDyxPO4 nanocrystals

Open Access Open Access

Abstract

We propose a novel material prepared by microwave-hydrothermal treatment, the tetragonal xenotime-type yttrium orthophosphate YPO4 nanocrystals doped by different concentrations of Dy3+. It may be suitable for laser-induced local heating of cancer tumors for hyperthermia. We heated a powder consisted of the nanoparticles by focused quasi-CW laser irradiation at different wavelengths in the near IR spectral range fitting the transparency window of biological tissues. The local temperature on the surface of the powder in the place of irradiation increases linearly with increasing laser power and increasing the Dy3+ concentration. At the same time the efficiency of local heating Φ = ΔT / (P f) (ΔT is a local temperature increase, f is an oscillator strength of absorption transition, and P is the quantity of laser power) is proportional to the energy of the initially excited electronic level. The proposed method allows for high rates of heating and cooling. The laser power used for heating was rather low, tens of milliwatts that together with short heating time to required temperature may result in extremely low doses of laser irradiation for heating.

© 2015 Optical Society of America

1. Introduction

We propose a novel material, the rare-earth doped nanoparticles (NPs), which may be suitable for photo-induced local heating of cancer tumors for hyperthermia [1]. The idea of heating is centered on the process of multiphonon relaxation (MR) of the optical excitation energy in the RE doped crystals. In the field of rare-earth doped fluorescent and laser materials intra-center multiphonon relaxation usually competes with radiative relaxation [2]. In the single frequency model of lattice vibrations a decrease of the phonon number p = ∆E/ℏωeff. bridging the energy gap ∆E between two electronic levels by one raises the rate of multiphonon transition by one or two orders of magnitude [3]. If the number of phonons is equal or less than three (p ≤ 3), the RE ions fluorescence almost completely quenches by multiphonon relaxation, because the rate of multiphonon transition is on the nanosecond or even picosecond time scale that is 105 – 107 times faster than the spontaneous emission decay rate of the RE ions.

However, the negative effect of MR in case of fluorescent materials can be used as a positive effect for nanoscaled heaters. We propose “non-fluorescent” nanocrystals instead of fluorescent ones. For this we reduce a fluorescent quantum yield from almost unity typical for metastable levels of the RE ions to 10−5 – 10−7 raising the rate of MR comparing to radiative rate. In so doing we choose the RE ion, which enables immediately after laser irradiation to start multistage (cascade) multiphonon relaxation process down to the ground level with efficiency of photon energy transformation to heat close to unity. A Dy3+ ion embedded into the YPO4 crystal matrix having simultaneously a wide phonon spectrum (ћωmax. = 1100 cm−1 [4]) and permiting up to 100% substitution of the Dy3+ ion for Y3+ [5] meets this requirement. Also, the phosphate water-dispersible crystalline nanoparticles can be produced rather easily as opposed to the oxide NPs, which are usually produced using solvent-free techniques or in oily solvents. These NPs are harder to disperse in water, which is essential for biomedical applications. Besides yttrium ortophosphate is biocompatible with human tissues. The energy level diagram of the Dy3+ ion allows for choosing laser excitation wavelength between 760 (6F3/2), 811 (6F5/2), or 914 nm (6F7/2) (Fig. 1and Fig. 2) in the near IR spectral range fitting the transparency window of biological tissues (750 – 950 nm [6]). It has real advantage over the fluoride crystal matrixes, where even in the LiYF4 host crystal with the most extended phonon spectrum among fluoride crystals the maximal phonon energy does not exceed ћωmax. = 560 cm−1 [7]. The use of the fluoride hosts would increase the number of phonons bridging the energy gaps for the 6H11/26H13/2 and 6H13/26H15/2 transitions of Dy3+ to four and five for LiYF4 and to six and seven for the LaF3 crystal host, respectively. The latter has the lowest maximal phonon frequency ћωmax. = 400 cm−1 [8] among fluoride crystals. As a result the MR rates of the 6H11/2 and 6H13/2 levels in the Dy3+ doped fluoride matrixes would have been comparable with their spontaneous emission decay rates, and significant amount of the optical excitation energy would have emitted by photons rather than phonons. However, for highly concentrated Dy3+ doped fluoride samples like DyF3 the concentration quenching may compensate the low phonon spectrum of the matrix.

 figure: Fig. 1

Fig. 1 Energy level diagram with indication of multiphonon transitions (solid arrows downward) in the Dy3+: YPO4 crystalline nanoparticles under direct laser excitation into the 6F3/2, 6F5/2, and 6F7/2 levels of the Dy3+ ion.

Download Full Size | PDF

 figure: Fig. 2

Fig. 2 Reflectance spectrum of the DyPO4 sample at room temperature measured by Laser Electronic LESA-01-Biospec spectra analyzer with tunable femtosecond Chameleon laser excitation. The spectrum is recorded by diffuse reflection of white light from a thin layer of powder. Y axis is the relative units.

Download Full Size | PDF

The heating kinetics for a nanoparticle suspended in a medium can be derived in the assumption that the thermal equilibrium in the phonon subsystem is set during the time τ much shorter than the flow time of radiative and non-radiative transitions, i.e. the lifetime of nonequilibrium phonons is much shorter than the flow times of radiative and non-radiative transitions

dT/dt=ab(Tθ)

The first term describes the temperature rise due to the absorption of laser radiation and transforming it into heat. The second term describes the loss of heat by transferring heat energy to the environment. We used Newton's law of cooling to obtain this term. Here θ is a temperature of environment medium. The first term is

a=(ND/CV)(1ηf)(τp/τ0)(σnano/σbulk)σif(ω)I(ω)dω,
where I(ω) is the spectral density of intensity of laser radiation; σif (ω) is an absorption cross-section of transition from an initial electronic state | i > to an excited state | f > in the same bulk crystal; ND is a number of rare-earth ions in the unit of the volume, i.e. their concentration; CV is a heat capacity at constant volume per unit volume; ηf is a fluorescence quantum yield of the excited electronic level; τp is a laser pulse width; τ0 is a repetition period of the pulses; σnano / σbulk is a factor taking into account the difference in the cross sections of nanocrystal and bulk crystal of the same compound [9];
b=Sh/VCV,
and S and V are a surface area and a volume of NP, respectively; h is a heat transfer coefficient. As a result we derived the equation for ΔT (t) = T - ϴ as

ΔT(t)=a[1exp(bt)]/b.

The main result is that the temperature of NP increases linearly with increasing the spectral density of absorbed laser radiation and concentration of RE dopant, and decreases with increasing the S/V ratio that is with the decreasing the size of nanoparticle. At the same time transformation of light to heat does not depend on the size of the NP (Eq. (2)). Therefore, the surrounding medium may be heated more for smaller nanoparticle.

Indeed, this theory does not claim to be an accurate quantitative description. It does not take into account the inhomogeneity of the intensity distribution of the laser beam, laser light scattering, heat transfer peculiarities in the powder, size distribution of the NPs, and a number of other factors, which have to be considered to construct a quantitative theory. To create one it is also necessary to obtain experimental data on the kinetics of the temperature distribution in the powder in the plane perpendicular to the propagation direction of the laser beam. These studies can be carried out later. However, the simplified theory can be used for qualitative interpretation of experimental regularities of nanoparticles powder heating kinetics. Employing the Y1-xDyxPO4 nanocrystals, we study the regularities of heating of the NPs under near IR laser irradiation.

2. Experimental results and discussion

Samples were prepared using microwave-hydrothermal treatment of freshly precipitated gels. The samples preparation and details on laser heating experiments are presented in the Appendix.

Laser heating by 811 and 914 nm wavelengths demonstrates almost linear dependence of the local temperature increase ∆T on the laser power of a powder surface hottest area (425 x 425 μm) taking from the central pixel of the camera image for direct excitation into the 6F5/2 and 6F7/2 levels of Dy3+ in the DyPO4 and Y0.525Dy0.475PO4 nanocrystals Fig. 3.Also, the heating increases in proportion with increasing of Dy3+ concentration. Qualitatively, these results are consistent with Eqs. (2) and (4).

 figure: Fig. 3

Fig. 3 The local temperature increase ∆T of the Y1-xDyxPO4 powder taken from the hottest pixel of the image of the powder surface on the IR camera after laser irradiation versus laser power. Excitation was done into the 6F5/2 (811 nm) and 6F7/2 (914 nm) levels of Dy3+ in the scanning microscope spot mode. Temperature measurements were done when approaching the steady-state. The laser parameters and details of the optical scheme of excitation are given in the Appendix.

Download Full Size | PDF

Also we found a linear increase of the local temperature increase ∆T of the powder hottest area with the increase in Dy3+ concentration in the range from 1 to 47.5 mol.% of Dy3+ and low decline from linearity for DyPO4 for all three excitation wavelengths, 760, 811, and 914 nm Fig. 4. We may attribute the latter to surface water evaporation.

 figure: Fig. 4

Fig. 4 The local temperature increase ∆T of the Y1-xDyxPO4 powder taken from the hottest pixel of the image of the powder surface on the IR camera after 100 mW laser irradiation versus the concentration of Dy3+. Excitation was done in the scanning microscope spot mode. Temperature meausrements were done when approaching the steady-state.

Download Full Size | PDF

We measured the heating efficiency of the powder Φ = ΔT / (P f) as a ratio of its local temperature increase (ΔT) to the product of the oscillator strength of the absorption transition (f) [10] and the quantity of laser power (P). We obtained the highest heating efficiency, more than one degree per mW when exciting into the top 6F3/2 level of Dy3+ in the DyPO4 nanocrystals (Table 1), and the lowest, less than 0.3 degree per mW, while exciting into the third top the 6F7/2 level, which both relax by one-phonon transition. Also, we excited the second top the 6F5/2 level, which relaxes by 2-phonon transition, and obtained higher heating efficiency than exciting into the third top the 6F7/2 level. We found that the heating efficiency does not correlate with the MR rate of the excited level. Otherwise it would be higher for the 6F7/2 level relaxing with the emission of just one phonon than for 6F5/2 relaxing with the emission of two phonons. We conclude that in the system under study the efficiency of heating is proportional to a number of multiphonon transitions with p ≤ 3 in the cascade nonradiative relaxation process, which is maximal for the upper 6F3/2 level (N = 12) and minimal for the lowest 6F7/2 level (N = 10), or in other words the higher is the initially excited level, the more electronic energy is transferred to heat, ceteris paribus. A decrease of the Dy3+ concentration reduces the heating efficiency (Table 1) in accordance with the dependence shown in Fig. 3.

Tables Icon

Table 1. The Heating Efficiency Φ of the Y1-xDyxPO4 Nanocrystals Depending on the Number (N) of Multiphonon Transitions with p ≤ 3 in the Cascade Nonradiative Relaxation Process (ℏωmax ≈1100 cm−1a)

At the same time, the heating time of the powder is very short. For example, the rise time of the temperature from room temperature to 340K at the hottest spot of the DyPO4 powder under pulsed 811 nm laser irradiation in the scanning microscope spot mode with average power of 30 mW is approximately one second only Fig. 5, upper curve. The temperature drops back to room temperature within the same time. This is an indication of inertialess heating, which enables setting precise duration for hyperthermia, which may be very important for selective treatment of cancer tumors without disturbing healthy tissues. Besides, the laser power used for heating was low, tens of mW only that together with fast heating time to the required temperature may result in low doses of laser irradiation of human body during hyperthermia treatment. We found that the rates of heating and cooling are independent on excitation wavelength Fig. 5. The laser irradiation at 850 nm, which is out of energy resonance with Dy3+ energy levels Fig. 2, does not indicate significant heating of the powder Fig. 5, lower curve, which confirms that the heating is a result of the multiphonon relaxation from the excited Dy3+ electronic states.

 figure: Fig. 5

Fig. 5 The local temperature kinetics of the hottest area of the DyPO4 nanocrystals powder (0.425 x 0.425 mm) taken from the hottest pixel of the image on the IR camera of the powder surface under laser irradiation in the scanning microscope spot mode with the average power of 30 mW at different excitation wavelengths, from top to bottom 811 (red), 914 (dark red), 760 (green), and 850 nm (blue). The laser was switched, on and off, every second.

Download Full Size | PDF

It is seen that the temperature of the NPs can be rather high, much higher than it is necessary for cancer hyperthermia treatment. However for colloidal solutions the maximal achievable local temperatures will be lower than for the powders. This requires a separate study including the development of direct and indirect methods for the temperature measurements.

We believe that it could be possible to synthesize the core-shell NPs with photon emissive core doped by Nd3+ ion [11] and heat emissive shell doped by Dy3+ using the same YPO4 crystal matrix for simultaneous near IR tumor imaging and cancer laser hyperthermia treatment.

3. Conclusion

We propose a novel material, the Y1-xDyxPO4 nanocrystals, which may be suitable for laser-induced local heating of cancer tumors for hyperthermia. The heating mechanism based on multiphonon relaxation of optical excitation. The local temperature of the nanoparticles powder linearly increases with increasing irradiation laser power and Dy3+ concentration. At the same time the efficiency of the local heating Φ is proportional to the energy of the initially excited electronic level. The proposed method allows for high rates of cooling and heating, which may result in precise spacing and timing during hyperthermia treatment. The laser power used for heating is rather low, tens of milliwatts that together with short heating time to the required temperature may result in low doses of laser irradiation of human body for treatment.

Appendix

1. Sample preparation

As starting compounds for preparation of the Y1-xDyxPO4 nanoparticles (x = 0.01, 0.05, 0. 475 and 1) we used DyCl3•6H2O (Aldrich, 99.995% purity), Y(NO3)3•4H2O (Aldrich, 99.999% purity) and K2HPO4•3H2O (Aldrich, 99.9% purity). For the synthesis we prepared solutions of 5 mmols of the mixture of DyCl3•6H2O and Y(NO3)3·4H2O, taken in corresponding stoichiometric proportions, in 10 ml of deionized water, as well as solution of 5 mmols of K2HPO4·3H2O in 30 ml of deionized water. After that we added the solution of rare-earth salts dropwise to the solution of phosphate under vigorous stirring and left it for 15 min keeping the stirring on. We diluted the freshly precipitated gel in mother solution with 10 ml of deionized water, transfered it into 100 ml Teflon autoclave and expose to microwave-hydrothermal (MW-HT) treatment (200°C, 2 hours) using a Berghof Speedwave-4 laboratory device (2.45 GHz, 1 kW maximum output power). After the treatment the samples were centrifuged, washed several times with deionized water and air-dried at 100°C for 5 hours.

We performed the X-ray diffraction analysis of the NPs using the «Rigaku D/MAX 2500» diffractometer (CuKα-radiation) and identified diffraction peaks using the JCPDS database. Instrumental broadening of the reflections was measured on the bases of standard material SRM-660 (LaB6). Physical broadening (β) of reflections was calculated using Voigt decomposition method. The mean size of the coherent scattering region (CSR) in the direction normal to the observed atomic planes was estimated using the Scherrer equation [12].

CSR=λ/(β×cos(θ)),
where λ is the wavelength of X-ray irradiation and θ is the Bragg angle for the particular reflection studied. The transmission electron microscopy (TEM) images of the samples were taken with Leo912 AB Omega microscope under accelerating voltage 100 kV.

2. Phase composition and morphology

XRD analysis of obtained samples (NPs powder) showed Fig. 6 that they consist of pure tetragonal phase with I41/amd space group, isostructural to xenotime YPO4. It is worthy to note that isostructural YPO4 and DyPO4 phases have almost equal lattice parameters, and due to lanthanide contraction [13] the radius of the Dy3+ ion is very close to the one of Y3+ ion. Therefore, regardless of the Y: Dy ratio in solid solution, the positions of maxima on XRD patterns remain almost the same. The synthesized samples of Y1-xDyxPO4 possess high degree of crystallinity. The mean size of the coherent scattering region (CSR) depends on the value of x and changes from 40±3 nm for x = 0.01 to around 150 nm for x = 1 (the precision of CSR size determination for large sizes is rather low due to small physical broadening).

 figure: Fig. 6

Fig. 6 XRD patterns of the Y1-xDyxPO4 nanoparticles with x = 0.01 (red) and 1 (black).

Download Full Size | PDF

Morphology of the synthesized nanoparticles was studied by means of TEM Figs. 7, Fig. 8, and Fig. 9. Nanoparticles are isotropic and rather uniform. Mean size of the particles determined using the TEM data is 40±15 nm for x = 0.01, 65±22 nm for x = 0.475, and 125±45 nm for x = 1, which is in good correlation with XRD results and confirms the high degree of crystallinity of all the samples. The width of the size distribution does not vary substantially for synthesized samples. It is close to normal with slight admixture of lognormal component, which decreases with the decrease in the dysprosium content. For nanoparticles of pure DyPO4 one can see Fig. 9 that apart from the main fraction of cuboid shape large particles, some amount of significantly smaller rod-like nanoparticles is present. Closer look to the inner structure of the larger particles and the remaining aggregates of the smaller nanoparticles allows one to suggest that formation of larger nanoparticles is due to oriented attachment and growth of rod-like nanoparticles, rather than due to Ostwald ripening. The same most likely applies to the Y0.525Dy0.475PO4 nanoparticles. As for the Y0.99Dy0.01PO4 nanoparticles it is hard to deduce from TEM micrographs whether they formed mainly by aggregation or growth.

 figure: Fig. 7

Fig. 7 TEM image and particles size distribution of the Y0.99Dy0.01PO4 nanoparticles.

Download Full Size | PDF

 figure: Fig. 8

Fig. 8 TEM image and particles size distribution of the Y0.525Dy0.475PO4 nanoparticles.

Download Full Size | PDF

 figure: Fig. 9

Fig. 9 TEM image and particles size distribution of the DyPO4 nanoparticles.

Download Full Size | PDF

3. Experimental details

We excited the Dy3+ ion directly into the 6F3/2, 6F5/2, or 6F7/2 levels by pulsed tunable femtosecond laser Coherent Chameleon Ultra II with 140 fs pulse duration, repetition frequency 80 MHz, and maximal average power up to 140 mW. We used 760, 811 and 914 nm wavelengths that is into the maxima of the spectral peaks of the 6H15/26F3/2, 6F5/2, and 6F7/2 optical transitions of Dy3+, respectively lying in the transparency window of biological tissues (750 – 950 nm). Approximately 65 mg of the nanoparticles powder was poured between two cover glasses with a width of 1.0 mm and a thickness of 0.2 mm, laid on a glass slide. The distance between the cover glasses amounts at 0.7 mm. So, the volume of the sample was 0.14 10−3 cm3 Fig. 10. The excess of the powder (powder protruding above the surface of the cover glass) has been removed from the surface. The laser beam was focused on the bottom surface of the powder layer through the slide. Temperature readout was taken from the upper layer of powder. The laser irradiation was done through scanning microscope ZEISS LSM 710 in a spot mode. The laser spot had the diameter of 10 μm. The ellipticity of the laser beam is 0.9 – 1.1. We measured the temperature of nanoparticles by Remote High Sensitive IR camera JADE MWIR SC7300M, Cedip sensitive in the mid IR (3 – 5 μm) spectral range with maximal time resolution 6.7 ms. The size of a single camera pixel was 530 x 530 μm, which is a limit for spatial resolution of the temperature measurements. The imageof the area in the spot mode was projected onto the four pixels of the IR camera.

 figure: Fig. 10

Fig. 10 Scheme of the experiment for temperature measurement.

Download Full Size | PDF

Acknowledgments

This work is supported by European Social Fund Mobilitas grant No. MTT 50.

References and links

1. D. Jaque, L. Martínez Maestro, B. del Rosal, P. Haro-Gonzalez, A. Benayas, J. L. Plaza, E. Martín Rodríguez, and J. García Solé, “Nanoparticles for photothermal therapies,” Nanoscale 6(16), 9494–9530 (2014). [CrossRef]   [PubMed]  

2. O. Svelto, Principles of Lasers (Springer, 1998).

3. Y. V. Orlovskii, R. J. Reeves, R. C. Powell, T. T. Basiev, and K. K. Pukhov, “Multiple-phonon nonradiative relaxation: Experimental rates in fluoride crystals doped with Er3+ and Nd3+ ions and a theoretical model,” Phys. Rev. B Condens. Matter 49(6), 3821–3830 (1994). [CrossRef]   [PubMed]  

4. A. A. Kaminskii, M. Bettinelli, A. Speghini, H. Rhee, H. J. Eichler, and G. Mariotto, “Tetragonal YPO4 – a novel SRS-active crystal,” Laser Phys. Lett. 5(5), 367–374 (2008). [CrossRef]  

5. E. N. Silva, A. P. Ayala, I. Guedes, C. W. A. Paschoal, R. L. Moreira, C.-K. Loong, and L. A. Boatner, “Vibrational spectra of monazite-type rare-earth orthophosphates,” Opt. Mater. 29(2-3), 224–230 (2006). [CrossRef]  

6. V. B. Loschenov, V. I. Konov, and A. M. Prokhorov, “Photodynamic therapy and fluorescence diagnostics,” Laser Phys. 10, 1188–1207 (2000).

7. S. A. Miller, H. E. Rast, and H. H. Caspers, “Lattice vibrations of LiYF4,” J. Chem. Phys. 52(8), 4172–4175 (1970). [CrossRef]  

8. R. P. Bauman and S. P. S. Porto, “Lattice vibration and structure of rare-earth fluorides,” Phys. Rev. 161(3), 842–847 (1967). [CrossRef]  

9. K. K. Pukhov, T. T. Basiev, and Yu. V. Orlovskii, “Spontaneous emission in dielectric nanoparticles,” JETP Lett. 88(1), 12–18 (2008). [CrossRef]  

10. R. Faoro, F. Moglia, M. Tonelli, N. Magnani, and E. Cavalli, “Energy levels and emission parameters of the Dy3+ ion doped into the YPO4 host lattice,” J. Phys. Condens. Matter 21(27), 275501 (2009). [CrossRef]   [PubMed]  

11. E. V. Samsonova, A. V. Popov, A. S. Vanetsev, K. Keevend, E. O. Orlovskaya, V. Kiisk, S. Lange, U. Joost, K. Kaldvee, U. Mäeorg, N. A. Glushkov, A. V. Ryabova, I. Sildos, V. V. Osiko, R. Steiner, V. B. Loschenov, and Y. V. Orlovskii, “An energy transfer kinetic probe for OH-quenchers in the Nd(3+):YPO4 nanocrystals suitable for imaging in the biological tissue transparency window,” Phys. Chem. Chem. Phys. 16(48), 26806–26815 (2014). [CrossRef]   [PubMed]  

12. B. D. Cullity and S. R. Stock, Elements of X-Ray Diffraction, (Prentice-Hall Inc., 2001).

13. A.F. Cotton, G. Wilkinson, Advanced Inorganic Chemistry (Wiley-Interscience, 1988).

Cited By

Optica participates in Crossref's Cited-By Linking service. Citing articles from Optica Publishing Group journals and other participating publishers are listed here.

Alert me when this article is cited.


Figures (10)

Fig. 1
Fig. 1 Energy level diagram with indication of multiphonon transitions (solid arrows downward) in the Dy3+: YPO4 crystalline nanoparticles under direct laser excitation into the 6F3/2, 6F5/2, and 6F7/2 levels of the Dy3+ ion.
Fig. 2
Fig. 2 Reflectance spectrum of the DyPO4 sample at room temperature measured by Laser Electronic LESA-01-Biospec spectra analyzer with tunable femtosecond Chameleon laser excitation. The spectrum is recorded by diffuse reflection of white light from a thin layer of powder. Y axis is the relative units.
Fig. 3
Fig. 3 The local temperature increase ∆T of the Y1-xDyxPO4 powder taken from the hottest pixel of the image of the powder surface on the IR camera after laser irradiation versus laser power. Excitation was done into the 6F5/2 (811 nm) and 6F7/2 (914 nm) levels of Dy3+ in the scanning microscope spot mode. Temperature measurements were done when approaching the steady-state. The laser parameters and details of the optical scheme of excitation are given in the Appendix.
Fig. 4
Fig. 4 The local temperature increase ∆T of the Y1-xDyxPO4 powder taken from the hottest pixel of the image of the powder surface on the IR camera after 100 mW laser irradiation versus the concentration of Dy3+. Excitation was done in the scanning microscope spot mode. Temperature meausrements were done when approaching the steady-state.
Fig. 5
Fig. 5 The local temperature kinetics of the hottest area of the DyPO4 nanocrystals powder (0.425 x 0.425 mm) taken from the hottest pixel of the image on the IR camera of the powder surface under laser irradiation in the scanning microscope spot mode with the average power of 30 mW at different excitation wavelengths, from top to bottom 811 (red), 914 (dark red), 760 (green), and 850 nm (blue). The laser was switched, on and off, every second.
Fig. 6
Fig. 6 XRD patterns of the Y1-xDyxPO4 nanoparticles with x = 0.01 (red) and 1 (black).
Fig. 7
Fig. 7 TEM image and particles size distribution of the Y0.99Dy0.01PO4 nanoparticles.
Fig. 8
Fig. 8 TEM image and particles size distribution of the Y0.525Dy0.475PO4 nanoparticles.
Fig. 9
Fig. 9 TEM image and particles size distribution of the DyPO4 nanoparticles.
Fig. 10
Fig. 10 Scheme of the experiment for temperature measurement.

Tables (1)

Tables Icon

Table 1 The Heating Efficiency Φ of the Y1-xDyxPO4 Nanocrystals Depending on the Number (N) of Multiphonon Transitions with p ≤ 3 in the Cascade Nonradiative Relaxation Process (ℏωmax ≈1100 cm−1a)

Equations (5)

Equations on this page are rendered with MathJax. Learn more.

dT/dt=ab(Tθ)
a=( N D / C V )(1 η f )( τ p / τ 0 )( σ nano / σ bulk ) σ if (ω)I(ω)dω ,
b=Sh/V C V ,
ΔT(t)=a[1exp(bt)]/b.
CSR = λ/(β×cos(θ)),
Select as filters


Select Topics Cancel
© Copyright 2024 | Optica Publishing Group. All rights reserved, including rights for text and data mining and training of artificial technologies or similar technologies.