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

Control of infrared cross-relaxation in LiNbO3:Tm3+ through high-pressure

Open Access Open Access

Abstract

The cross-relaxation process between the lower excited states of thulium ions has a strong influence on its main infrared emissions, but also in the population of higher excited states that lead to characteristic blue upconversion. This work investigates this process in LiNbO3:Tm3+ by means of time-resolved spectroscopy at high pressure. It is demonstrated that through the application of high-pressure it is possible to enhance its probability and to investigate its influence on the photoluminescence spectra and corresponding lifetimes of Tm3+. The results are analyzed in terms of the effect of high pressure on parameters such as Tm3+-Tm3+ distance through the equation-of-state of LiNbO3, refractive index or Tm3+-Tm3+ energy transfer characteristics (absorption/emission overlap integral), to conclude that the major multipole interaction responsible for cross-relaxation is the quadrupole-quadrupole interaction. This conclusion supports and clarifies previous dynamical models for energy transfer on the basis of spectroscopic studies carried out in LiNbO3:Tm3+ as a function of Tm3+ concentration.

© 2015 Optical Society of America

1. Introduction

Tm3+-activated materials are extensively investigated because of their luminescent properties, which are important in two aspects: due to a broad infrared luminescence emission around 1.8 µm (3F43H6 transition), a relevant wavelength in fiber-based telecommunications [15], and due to their capabilities as ultraviolet and blue emitter (at 371 nm, 476 nm) following infrared (IR) excitation, i.e. through an up-conversion process, that makes these materials particularly interesting in nanomedicine and bioimaging [6, 7]. Besides, LiNbO3:Tm3+combines the optical capabilities of Tm3+ with the electro-optic, acousto-optic and non-linear properties of LiNbO3 making it an attractive material to develop optoelectronic devices. As a consequence of the outstanding properties of LiNbO3:Tm3+, it has been recently used as active material for photon-echo quantum memories [8, 9] and laser action has been investigated and demonstrated in several configurations [1, 2], including π-polarized cw laser oscillation of Tm3+ ions in Zn-diffused channel waveguides operating at 1.76 µm under 795 nm pumping (to 3H4 state of thulium) [10]. Nevertheless, the efficiency slope reported for these waveguide lasers is lower than that reported for other Tm3+ lasers [11, 12], a result that has been attributed to either a non-optimized cavity design or losses due energy transfer processes that affect Tm3+ luminescence.

This commonly used excitation wavelength is linked to one spectroscopic characteristic that makes Tm3+-doped materials outstanding for optical applications: the occurrence of an infrared cross-relaxation (CR) mechanism (3H43F4:3H63F4) that yields photoluminescence quenching of the near-IR emission at 795 nm, while doubly populates the 1.8 µm emitting level (3F43H6 transition). The knowledge of the depopulation mechanisms from 3H4 state and of how structural changes in the host affect it, result crucial to unveil the efficiency of Tm3+ lasers and other characteristic optical properties of Tm3+-doped materials.

It is obviously possible to alter the probability of energy-transfer by doping with different concentrations of Tm3+ ions, since the distance between donor and acceptor ions in the host is different for different concentrations [13]. However, an increase in doping concentration introduces as well charge compensation defects in the lattice, and thus modifies the local environment of the ions. An alternative approach is the application of external hydrostatic pressure. Although applied pressure only produces a small variation in the distance between donor and acceptor for any practical range of pressures, it offers the advantage of being a highly controlled perturbation that affects the whole range of parameters influencing ion-ion interaction. Consequently, it offers a very rigorous test of energy transfer theoretical models.

Particularly interesting in this case, is the effect of pressure on the transition frequencies of Tm3+. It must be pointed out that the above mentioned CR scheme is a non-resonant process [13, 14], and therefore, the energy mismatch between the emission of the sensitizer and the absorption of the acceptor has to be energetically bridged by involving host phonons. Generally, the larger the mismatch is, the lower the efficiency of the CR process. In this way, it is expected that the application of high pressure increases the crystal-field strength on Tm3+ dopants producing a shift of the energy states, thus modifying the energy mismatch and CR probability, while keeping unaltered the chemical composition of the material.

The aim of this work is to study the energy shift of Tm3+ states and their dynamics under high-pressure conditions, in order to understand its effect over the infrared CR and also to establish the most advantageous conditions to enhance CR efficiency.

2. Materials and methods

The LiNbO3:Tm3+ sample studied in this work has been grown by the Czochralski technique in air atmosphere using a platinum crucible and automatic diameter control by crucible weighting. The starting raw materials were congruent LiNbO3 ([Li]/[Nb] = 0.945) and thulium oxide (purity 99.99%), both of them are commercially available and were used without further purification. An optically selected sample from the as-grown boule was analyzed through x-ray diffraction to check the crystal structure and to verify the [Li]/[Nb] ratio. The [Tm3+] concentration was determined both by x-ray fluorescence and optical absorption to conclude a value [Tm3+] = 2.5 mol% [15].

The absorption and emission/excitation spectra at ambient conditions were obtained with a spectrophotometer (Cary 6000i, Varian) and a fluorometer (Jobin-Yvon FluoroMax-2), respectively. For blue up-conversion luminescence experiments, the sample was excited with a Ti-Sapphire laser operating at 795 nm (100 mW).

Hydrostatic pressure experiments in the 0-12 GPa range were carried out on a membrane-type diamond anvil cell (DAC) and a Meryl-Basset-type MALTA DAC. Inconel gaskets of 200 μm thickness were pre-indented down to 40 μm thickness. Suitable 150 μm-diameter holes were drilled with a Betsa motorized electrical discharge machine. The DAC was loaded with suitable LiNbO3:Tm3+ single crystal and ruby microspheres (<10 µm diameter) using either methanol-ethanol-water mixture (16:3:1) or silicon oil as pressure transmitters. In all cases, experiments were performed in nearly-hydrostatic regime and the pressure was calibrated from the ruby photoluminescence [16, 17]. For cw excitation inside the DAC, the 476.5 nm line of a Krypton laser (Coherent CR-500K) has been used, while for pulsed excitation, the tunable beam of an OPO (Vibrant B 355 II) has been focalized backward on the sample with a 20 × microscope objective. The emission was collected with a microscope objective that allows the selection of the measured signal (sample or ruby probes). For lifetimes and time-resolved emission and excitation spectra with pulsed excitation, the emission light was dispersed by a 0.32 m-focal monochromator (Horiba-Jobin-Yvon Triax HR 320) and detected with a fast intensified charge-coupled device (iCCD) with 5 ns time resolution. The excitation spectra were corrected from the pulsed energy using a beam-splitter together with a power-meter. For lifetime measurements, the photoluminescence decay signal was recorded with the iCCD. In continuous mode excitation, the fluorescence signal was carried out by optical fiber for spectral analysis in the 600 – 1700 nm range using Ocean spectrometers (HR2000 + and NIRQuest512) equipped with Si and InGaAs array detectors, respectively, using a prototype microscope [18].

3. Infrared cross-relaxation mechanism

Tm3+ ions have a rich energy level structure arising from their 4f12 electronic configuration. Some of these levels are equally spaced, allowing energy transfer between Tm3+ ions. Particularly, a near-infrared CR process, that can be found in a wide variety of hosts [14, 1921], takes place between the 3H4 excited state and the ground state (3H6) and populates the intermediate state 3F4 (3H43F4:3H63F4) following the scheme of Fig. 1 (right). The activation of this mechanism yields a reduction of the 3H43H6 infrared emission (795 nm) and an increase of the population of 3F4 state, and thus it eventually produces an intensity enhancement around 1700 nm (3F43H6). This CR process has important consequences also for the upper excited states of Tm3+, directly responsible for ultraviolet (1D23H6 at 371 nm) and blue (1G43H6 at 476 nm) light emissions, since they can be populated through upconversion mechanisms. These mechanisms start on 3H4 and 3F4 infrared excited states, and thus the CR process that links these two states and modify their populations affects the intensity and dynamics of the ultraviolet and blue emissions [2225]. A complete characterization of these processes implies thus to study all involved emissions that can be relevant or affect CR.

 figure: Fig. 1

Fig. 1 (A) Optical absorption and (B) up-conversion photoluminescence via excitation at 795 nm of LiNbO3:Tm3+ at ambient conditions. On the right a partial scheme of Tm3+ energy states and the selected excitation scheme is added.

Download Full Size | PDF

In this work we choose to directly populate the excited state 1G4 with 476 nm laser excitation, a strategy that allows the study of all the emissions and associated dynamics arising from this and lower excited states. This excitation provides the most efficient pumping for 1G4 multiplet (Fig. 1) and, subsequently, after relaxation via radiative and/or non-radiative paths, also populates the 3H4 level, from which the cooperative energy transfer mechanism under study starts. In addition, this 476 nm excitation scheme is advantageous to characterize radiative emissions associated with 3H4 state, since it is far from all the emission bands of interest and thus avoids unwanted overlap with the excitation beam.

With the aim of analysing the effect of external pressure in energy transfer efficiency, a single crystal of LiNbO3:Tm3+ (90 × 80 × 20 μm3) was loaded in a membrane DAC for time-resolved spectroscopy. The selected pulsed-excitation wavelength (476 nm at ambient pressure) was finely re-tuned at every pressure to ensure the maximum emission intensity.

Incidentally, under the selected excitation wavelength, other Tm3+ luminescence emissions arising from some intermediate levels that also become sufficiently populated can also be monitored. All these emissions can affect not only CR processes in LiNbO3:Tm3+ (altering other excited states populations of levels participating in the CR path, such as 3F4, for instance) but also other important optical processes observed in the material. In fact, the excitation scheme adopted here is somehow the reversal of multiphoton up-conversion processes that lead to blue luminescence via infrared excitation at 795 nm (Fig. 1(b)), and are observed in LiNbO3 and other Tm3+-doped materials of interest. Therefore we shall present a complete characterization of pressure dependence of all the luminescence emissions observed under 476 nm pumping.

3.1. Pressure dependence of the peak position and intensity

The pressure dependence of the emission spectrum associated to all the observable transitions from 600 nm up to 1700 nm has been investigated under both continuous and pulsed excitation, and from ambient pressure (P0) up to 12 GPa. To ensure suitable noise-to-signal ratios, different integration times have been considered. Figure 2(a) shows the pressure dependence of the time-resolved emission spectra associated with the main emissions from 3H4 state (3H43H6) as well as 1G43H5. In the figure, vertical dashed lines have been added to indicate the position of the main peak of each transition at the starting pressure. The observed variations indicate that high pressure triggers red shift and broadening of the different peaks. Furthermore, the overall peak intensity strongly decreases with pressure, which could be connected with the activation of non-radiative mechanisms and CR processes. In particular, for 3H43H6 this modification would be related to a decrease of the photoluminescence lifetime, as will be discussed later on.

 figure: Fig. 2

Fig. 2 (A) Pressure dependence of the emission spectrum corresponding to 3H43H6 and 1G43H5 transitions after 476 nm excitation. (B) Decomposition of the emission spectrum at 2.2 GPa into four Lorentzian peaks. The inset shows the variation of the emission peaks with pressure.

Download Full Size | PDF

Each peak in the emission band can be associated to transitions from different Stark levels of LiNbO3:Tm3+ as assigned elsewhere [22]. Following this assignment, to analyze the intensity, position and widths of the different peaks, each spectrum has been fitted to four Lorentzian profiles. Figure 2(b) shows an example of the resulting fit and the variation of the position of the emission peaks with pressure is summarized in the figure inset.

The pressure dependence of the overall visible-infrared emission spectrum has been studied under cw excitation at 476.5 nm, with the aim of exploring changes induced in the emission spectra under the same excitation/pressure conditions (Fig. 3). The main emission peaks shown in the figure have been analyzed following the same procedure used in Fig. 2(b). The results obtained for the shift of the different peaks are given in Table 1.It must be noted that the intensity of 3H43H6 peak (not shown in Fig. 3 since it is in the detection limit of the spectrometer and has been analyzed separately in Fig. 2) decreases with pressure, but its variation is slower than that observed in the time-resolved spectra of Fig. 2(a). This difference in the intensity pressure dependence is entirely due to pressure-induced excitation detuning when the excitation is achieved at the fixed wavelength of 476.5 nm, and not retuned at every pressure as in Fig. 2. In fact, taking into account that the red-shift of the 3H61G4 absorption peak at 476 nm (Fig. 2(a)), derived from the 3H43H6 and 1G43H4 pressure shifts, is about −2 cm−1/GPa and the FWHM is 100 cm−1, the pumping capability at 476.5 nm increases by about 70% from 0 to 12 GPa due to pressure tuning enhancement of the absorption band. Therefore, the associated increase in 1G4 population, and correspondingly of 3H4 state with increasing pressure, palliates the loss of emission intensity due to non-radiative de-excitation. Consequently, as mentioned before, the pressure-induced intensity decrease is less rapid here than in Fig. 2(a).

 figure: Fig. 3

Fig. 3 (A) Variation of the emission spectra of LiNbO3:Tm3+ (2.5 mol%) in the visible and infrared ranges upon cw excitation at 476.5 nm. The infrared 3H43H6 emission peak at 12579 cm−1 (795.0 nm) shown in Fig. 2(a) is missed here because it appears at the spectral limit of the visible spectrometer. (B) Magnification of the variation with pressure of the 3F43H6 and 1G43F3 emission bands, and (C) pressure shifts of levels involved in the CR (3H63F4: 3H43F4).

Download Full Size | PDF

Tables Icon

Table 1. Peak position at ambient pressure, ωi(P0), and linear pressure shifts, (∂ωi/∂P), of the emission peaks associated to all the measured transitions. The most intense ones are highlighted in bold type.

Different Stark sublevels experience different shifts under pressure, as shown in Table 1. Particularly interesting, is the case of 3H4 and 3F4 states, since they are involved in the CR process we are aiming to analyze. There is a noteworthy pressure-induced blue shift undergone by 3F43H6 peak ( +1.6 cm−1/GPa) and a red shift of the most intense energy peak linked to 3H4 state (3H43H6, −2.1 cm−1/GPa). As a consequence, the corresponding 3H4 and 3F4 sublevels reduce significantly their energy distance (see Fig. 3(c)) at higher pressures, which would imply a strong red shift for the 3H43F4 transition (−3.7 cm−1/GPa) related to the same Stark sublevels that gave rise to the same exact 3H4 and 3F4 transitions commented before (It must be noted that this 3H4 transition is overlapped with 1G43F3 and cannot be accurately measured). The opposite shifts of 3F43H6 and 3H43F4 transitions, directly involved in the CR process (see Fig. 1), reduce substantially the energy mismatch between both transitions, thus enhancing CR probability as pressure increases. In this way the CR energy-mismatch is 509 cm−1 at P0 and reduces to 445 cm−1 at 12 GPa; i.e. there is a reduction of ca. 12%. This result is crucial to explain the CR enhancement under high-pressure conditions yielding reduction of blue and infrared emission intensities.

Finally, the relative intensity variations of the main emission peaks with pressure are presented in Fig. 4. For that purpose, the intensities have been normalized taking the 3H43H6 emission band at 795 nm as reference. Comparing the variation of the relative intensities of Fig. 4 with the pressure dependence of the 3H43H6 absolute intensity shown in Fig. 2(a), it can be concluded that all the emissions reduce their intensity with pressure, although each transition shows a different decrease rate. It must be noted that under cw excitation, the 3H43H6 peak experiences the strongest intensity reduction with respect to 1G43H4, 3H43F4, and 3F43H6 peaks, whose relative intensities are reduced less than about 50% of 3H43H6 peak intensity. This result is important to interpret and justify the CR processes as the main de-excitation path responsible for the loss of intensity with pressure.

 figure: Fig. 4

Fig. 4 Variation of the relative intensity of the main emission peaks of LiNbO3:Tm3+ with pressure. The relative variation for each peak is compared with the variation of the 3H43H6 peak taken as reference.

Download Full Size | PDF

3.2. Pressure dependence of the emission lifetime

Generally, energy transfer processes shorten the lifetime of the donor state, in our case 3H4. Eventually, it is possible that the acceptor state (here 3F4) accounts as well for some modification, since the energy transfer population path can be active at the same time the acceptor state is being depopulated, producing some lengthening of the decay. However, in our case the acceptor state 3F4 accounts for a lifetime more than one order of magnitude longer than the donor state, and thus no effect is expected on its decay time [14]. Consequently, the following study will be focused only on 3H4 lifetimes.

The temporal dependence I(t) of the emission intensity of the 795 nm emission associated to 3H43H6 transition has been studied as a function of pressure in the range 0 - 10 GPa. Figure 5(a) depicts the temporal evolution, in a semilogarithmic scale, of the intensity at three particular pressure values. As it can be observed I(t) behaves as a single exponential in the explored pressure range, and its lifetime decreases from 50 μs at ambient pressure to 34 μs at 10 GPa, following an approximate linear dependence with pressure (Fig. 5(b)).

 figure: Fig. 5

Fig. 5 (A) Intensity decay curves corresponding to the 3H43H6 emission (795 nm) after pulsed excitation into the1G4 state (475 nm) in LiNbO3:Tm3+ (2.5 mol%) at different pressures. The decays can be fitted to single exponential functions (blue solid lines). (B) Pressure dependence of the associated lifetime calculated by fitting the experimental data I(t) to a single-exponential function. The grey dotted line is a linear fit to the data, added to guide the eye.

Download Full Size | PDF

All the experimental results presented above show the strong influence of high pressure on the spectroscopic properties of Tm3+ ions in LiNbO3. Particularly, the intensity of 3H43H6 emission at 795 nm, especially in comparison to 3F43H6, and its lifetime behaviors indicate a strong dependence of the emission quantum efficiency. Consequently, the characteristic infrared CR process of Tm3+is behind the pressure induced changes, becoming, possibly, one of the main de-excitation paths of 3H4 state.

In general, the pressure-induced intensity reduction, as well as the lifetime shortening, can be related to different pressure dependent effects, such as variations in the phonon energy and/or density of states, energy gaps, oscillator strengths or energy transfer probabilities. Bearing in mind all these effects, we have analyzed the different contributions to the relaxation of 3H4 state. Therefore, the main objective of the present section is to quantify the possible incidence of these different variables, how they depend on pressure and to unveil which one is the most relevant.

There are two major ways in which pressure can influence the de-excitation transition probability, i.e. the reciprocal lifetime. On one hand by modifying the single-ion experimental lifetime (limit of low doping concentrations), either through radiative probability, 1/τrad, or through non-radiative multiphonon probability, 1/τNR, on the other hand by modifying the energy-transfer CR probability, WET. In terms of transition probability, it can be written as:

1τexp=1τrad+1τNR+WET=1τW0+WET
where τexp is the experimental lifetime of the donor state, 1/τrad and 1/τNR are the radiative and non-radiative multiphonon relaxation probabilities, respectively. In the right-hand side, the parameter τW0 is the experimental lifetime at low doping concentration, where energy transfer is absent (WET ≈0). Considering Eq. (1), the experimental lifetime associated with the donor level 3H4 seems to be an adequate parameter to analyze the changes produced by pressure in 3H4 state characteristics and in the infrared CR: 3H43F4:3F43H6.

The possibility of an increase of 1/τrad as main effect for the experimental lifetime decrease is ruled out, since this would represent an increase in transition probability and therefore an increase of the emission intensity, contrary to observations (Fig. 2(a)). Also the effect of pressure over τNR, which is analyzed in the next section, is negligible. Finally, we must consider that if the effect of high pressure on WET makes it larger, it would also make lifetime shorter, in agreement with experimental observations.

(a) Multiphonon relaxation process: WNR

One of the simplest options to account for 3H4 lifetime reduction is an increase in the multiphonon relaxation probability (Eq. (1)). It is known that pressure modifies the phonon frequency and density of states, an effect that has been extensively studied in LiNbO3 [26, 27]. For lanthanide ions, it is well established that the probability of multiphonon relaxation (WNR) between two states depends exponentially on the energy gap between both levels (ΔE), following the well-known “energy gap law” [2830]:

WNR=βelexp[α(ΔE2ω)]
where ħω is the energy of the highest energy phonon, and βel and α are constants only depending on the host and not on the dopant ion. In the case of LiNbO3:Tm3+, the experimental non-radiative probability of 3H4 state for low-dopant concentration of Tm3+ is WNR = 0.5 × 10−3 µs−1 [31]. This value fits fairly well to the estimates from the energy-gap law taking the α and βel parameters for LiNbO3: α = 3.4 × 10−3 cm and βel = 5.25 × 106 s−1 [32], the energy gap corresponding to the closest low-lying excited state 3H5: ∆E(3H4 - 3H5) = 4337 cm−1 (Table 1, Fig. 1), and the highest energy phonon, ħΩ = 880 cm−1 at ambient conditions [33].

In Eq. (2) the more sensitive parameters to pressure are ħω and ΔE. For LiNbO3, ħω shifts with pressure at a rate of 3.3 cm−1/GPa [27], while the energy gap (3H4 - 3H5) increases with pressure at a rate of 2.2 cm−1/GPa (Table 1). Bearing these values in mind, the increase in the multiphonon probability with pressure can be easily calculated through Eq. (2). We obtain a value of WNR = 8.2 × 10−4 µs−1 at ambient pressure and 9.7 × 10−4 µs−1 at 11 GPa. This implies a total variation, ∆WNR = 1.5 × 10−4 µs−1, which is an order of magnitude smaller than the total CR transfer probability experimentally observed, ∆WET = 10.5 × 10−3 µs−1 (Fig. 6(a)). In fact this non-radiative probability contribution (1% of the total variation) is within the experimental accuracy, and thus can be considered as a negligible contribution to data of Fig. 5.

 figure: Fig. 6

Fig. 6 Dependence of the measured transfer probability on the hydrostatic pressure. The experimental WET data (blue points), have been obtained from the measured lifetime using Eq. (1), taking τW0 = 230 μs. The blue dashed-dotted line has been drawn to guide the eye. The figure includes the probabilities, calculated from Eqs. (3) and (5), for the three main orders of the interaction (d-d, d-q and q-q). (A) The calculated probabilities consider only variations in RDA due to pressure-induced volume compression by means of the LiNbO3 equation of state. (B) Calculated energy-transfer probabilities due to both variations in RDA and refractive index of LiNbO3.

Download Full Size | PDF

(b) Energy transfer mechanism: WET

Energy transfer processes, apart from triggering intensity changes on the emissions of the levels involved in the mechanism, also influence the lifetime of the donor level (3H4 in the present case) since it is an additional relaxation path for that excited state (Eq. (1)). Considering the mentioned changes in the intensity of the emission bands (Fig. 3), the results obtained from the lifetime evolution of 3H4 state (Fig. 5) together with the irrelevance of non-radiative multiphonon relaxation processes analyzed previously, it turns out that energy transfer should eventually increase the transition probability.

In order to make quantitative estimations of WET(P), it is useful to look at the microscopic (ion to ion) definition of WET. Within a multipolar expansion of the ion-ion interaction, WET is a sum of different contributions that can be written as [34, 35]:

WET=CDA(S)RDAS
where RDA is the donor-acceptor distance; S is a parameter that changes with the order of the interaction, being 6 for dipole-dipole (dd), 8 for dipole-quadrupole (dq), 10 for quadrupole-quadrupole (qq), etc.; and CDA(S) is called energy-transfer microparameter and is defined by a different expression depending on the order of the interaction [35, 36].

The strong dependence of WET on the donor-acceptor distance makes it very sensitive to variations in doping concentration or external pressure. This behavior is consistent with the observed lifetime variation as a function of Tm3+ concentration, reported at ambient pressure elsewhere [14]. The lifetime reduction from τexp = 230 μs for [Tm3+] = 0.06 mol% to τexp = 50 μs for [Tm3+] = 2.5 mol% observed experimentally in LiNbO3:Tm3+ is ascribed to de-excitation via energy-transfer CR, thus showing the relevance of this mechanism.

It has been already demonstrated in several hosts that in an accurate description of the CR process between Tm3+ ions 3H43F4:3H63F4, the dipole-dipole order is unable to describe it, and thus dipole-quadrupole and quadrupole-quadrupole orders must be taken into consideration. Particularly, the quadrupole-quadrupole order interaction is very relevant since it usually provides the best CR description as a function of the [Tm3+] concentration (i.e. RDA) [31, 3740]. The microparameters corresponding to different order parameters CDA(S) are defined as [35, 36]:

CDA(6)=34c4QA4πn4τ0LD(E)LA(E)E4dE;CDA(8)=135α6c6QA4πn4τ0gDgA*gAgD*LD(E)LA(E)E6dE;CDA(10)=225ε8c8QA4πn4τ0LD(E)LA(E)E8dE
where c is the speed of light; n the refractive index at the wavelength of absorption/emission; QA = σabs(E)dE, the integrated absorption cross section of the transition; τ0 the radiative lifetime; the functions LD(E) and LA(E) in the overlap integrals are the normalized peak functions of the donor (emission) and acceptor (absorption), respectively; and gD and gA are the initial state degeneracy of donor and acceptor (asterisk refers to final states).

Looking at Eqs. (3) and (4), it turns out that WET will be sensitive to pressure through different magnitudes. For instance, pressure compresses the material leading to a reduction in the donor-acceptor distance (RDA). Consequently, the energy transfer probability would be larger at higher pressures as far as the RDA dependence is concerned. We can quantify this effect using the Murnaghan’s equation of state for LiNbO3:

VV0=(RR0)3=(1+B0'(PB0))1B0'
where the zero-pressure bulk modulus and its pressure derivative for LiNbO3 are B0 = 134 GPa and B0' = 2.9, respectively [41]. R/R0 refers to the relative distance reduction at a given pressure, and it applies to the averaged lattice parameters and, in general, any interatomic distance like RDA.

If we now rewrite Eq. (3) as WET = W0(R0/R)S, with W0 = CDA/(R0)S, and replace the distance dependence by the Murnaghan equation of state, it is easy to calculate how the pressure will modify the energy transfer probability due to RDA shortening (which is about −20% change for a pressure of 10 GPa). At this point W0 is considered as a constant, and is selected to allow every curve to have the WET value found experimentally at ambient pressure. For comparison purposes, the experimental values, expressed as an energy transfer probability, have been derived from τexp through Eq. (1) using τW0 = 230 μs [14]. The results for the three main orders of the interaction (S = 6, 8 and 10) are shown in Fig. 6(a).

It can be deduced from Fig. 6(a) that the shortening of RDA has a relevant effect in the pressure dependence of energy transfer probability, and as expected from Eq. (3) this effect is stronger for higher orders of the interaction. As previously mentioned, the quadrupole-quadrupole interaction provides the best description of this specific CR process between Tm3+ ions. However, even though its contribution is larger than the dipole-dipole and dipole-quadrupole ones, it is not enough to explain the experimental energy-transfer probability shown in Fig. 6(a). Although higher orders of the interaction could be added to explain the observed WET(P) variation, this option is not consistent given that the probability of such orders is negligible [31, 37, 38, 40].

Next we focus on the energy-transfer microparameter defined in Eq. (4), the pressure dependence of which was assumed to be constant in our previous analysis. It can be seen that in every case, its value is dependent to the fourth power of the refractive index. Due to the elasto-optic effect in LiNbO3, the refractive index increases with pressure, and its variation can be estimated through the differential Clausius-Mossoti equation by:

ΔVV=6n(n21)(n2+2)Δn
The relative volume variation can be transformed to pressure through the equation-of-state (Eq. (5)), hence the variations of the refractive index can be now introduced into Eq. (4) to calculate WET(P). Although this equation is deduced for an isotropic material, it can be used for an anisotropic system like LiNbO3 considering the averaged refractive index. For illustrating purposes, the refractive index of LiNbO3 at the overlap wavelength of the CR process increases from n = 2.225 at ambient pressure [42] to n = 2.359 at 10 GPa, what implies a relative contribution to WET of 21%. Following this idea, the effect of the modifications in the refractive index on WET has been calculated for the three orders of interaction. The results are shown in Fig. 6(b). It is clear that, as it could be expected from Eq. (4), an increase of the refractive index provokes a reduction of the energy transfer probability. This reduction appears to be stronger than the positive effect of the reduced distance between ions, since now the total calculated probability of CR is evolving towards lower values for higher pressures. Therefore, additional effects have to be considered to successfully explain the experimentally measured probability increase.

The last pressure-sensitive magnitude affecting microparameters in Eq. (4) is the overlap integral, since the pressure-induced shifts of the donor-emission and acceptor-absorption bands can substantially modify the energy transfer probability. According to the results shown in Table 1, donor emission and acceptor absorption approach each other at higher pressures (Fig. 3(c)), yielding increase of the overlap integral, and thus of the energy-transfer probability. It must be observed (Fig. 7(a)) that because the CR process is non-resonant, the overlap mainly takes place in a small energy range around the tails of the involved bands. By considering their respective peak shifts of −3.7 cm−1 (6544 cm−1) and +1.6 cm−1 (6035 cm−1) (see Fig. 2(c) and Table 1), the variation of the overlap integral with pressure can be calculated on the basis of fitted Lorentzian line-shapes that reproduce the spectra at atmospheric pressure in the overlap range. Then, the obtained line shapes of the donor and acceptor are normalized following the standard procedure to calculate the spectral overlap [35]. It has to be mentioned that the shift of the phonon energy can be considered as well in the overlap, since as it has been mentioned, the process is non-resonant and it could slightly modify the cross-relaxation probability. Nevertheless, as previously stated, the effect of pressure on the energy of the phonons is mostly negligible compared to other contributions, and no differences have been observed introducing this correction to the calculations.

 figure: Fig. 7

Fig. 7 (A) Spectral overlap of the donor emission and acceptor absorption in the considered cross-relaxation process. (B) Experimental WET (blue points) and calculated (orange dotted line) predictions for quadrupole–quadrupole interaction considering all possible contributions: RDA variations, refractive index changes and spectral overlap modification. The grey dashed lines have been calculated considering a ±0.1 cm−1/GPa error estimation in the shift coefficients of the peaks in order to illustrate the uncertainty of the calculation.

Download Full Size | PDF

The effect of variation of the overlap integral together with contributions from RDA and refractive index to the energy-transfer probability is shown in Fig. 7(b). The results clearly indicate that the overlap integral contribution definitively accounts for the experimental energy-transfer probability WET(P). The calculated WET fits fairly well to the experimental data by considering mainly the quadrupole-quadrupole (S = 10) interaction (orange solid line in the figure), while other interaction orders (d-d or d-q, not shown in the figure for clarity) give results substantially lower than the experimental values. This conclusion is in agreement with previous studies in which CR is analyzed as a function of Tm3+ concentration [37, 38, 40].

As a further analysis to understand the accuracy of the calculations, we have estimated the outcome that a ± 0.1 cm−1/GPa experimental error in each peak shift would have on the final result. It is clear that an error in the calculated shift would only affect to the contribution coming from the overlap integral, and that it would affect twice, since two different peaks are involved (Fig. 7(a)). The obtained result, shown between dashed lines for the quadrupole-quadrupole interaction in Fig. 7(b), accounts for the observed data within the experimental accuracy of the experiments.

Present results provide fully consistency on the microscopic mechanisms describing the CR process in LiNbO3:Tm3+ by analyzing the effects of volume reduction but keeping constant the Tm3+ content. For this purpose the use of high-pressure spectroscopy is crucial to achieve this goal. This demonstrates that the main reason for the reduction of 3H4 lifetime and the concomitant loss of emission intensity at 795 nm when pressure increases, is caused by an increase of the energy-transfer CR probability through three main combined mechanisms: reduction of donor-acceptor distances, increase of refractive index and increase of the overlap integral between donor emission and acceptor absorption bands.

Up to that point, W0 has been used as a fitting parameter at zero-pressure to separately study the influence of each magnitude on the total WET. However, quantitative estimations on the basis of quadrupole-quadrupole interaction using W0 = CDA/(R0)S and Eqs. (3) and (4), enable us to ensure that the quadrupole-quadrupole interaction is mainly responsible for CR. In fact, the experimental value WET(0) = 1.5 × 10−2 µs−1 (Fig. 7) can be accounted for within this model using Kushida’s equations for energy transfer, S = 10 and R0 = 8.4 Å [43]. So that we obtain W0 = 1.5 × 10−2 μs−1. The employed R0 value deviates only 5% of the actual value 8.0 Å derived on the assumption of a homogeneous distribution of Tm3+ (2.5 mol%) in LiNbO3, and thus confirms the adequacy of quadrupole-quadrupole interaction as most likely to CR. Instead, the next term contributing to the probability, i.e. the dipole-quadrupole (S = 8) is, according to Eqs. (3) and (4), two orders of magnitude smaller than the quadrupole-quadrupole interaction (S = 10) using R0 = 8.4 Å, thus ruling out this order interaction in CR.

4. Conclusions

We have demonstrated that the calculated energy-transfer probability fits fairly well to the experimental data by considering quadrupole-quadrupole interaction. This conclusion is in agreement with previous works in which CR is analyzed as a function of the Tm3+ concentration. The present work provides fully consistency on the microscopic mechanisms describing the CR in LiNbO3:Tm3+ by analyzing the effects of volume reduction while keeping constant the Tm3+ content. We also show that the main reason for the strong reduction of the main 795nm (3H43H6) emission intensity of Tm3+ and the corresponding lifetime decrease with pressure is caused by an increase of the energy-transfer CR probability through three combined mechanisms: reduction of donor-acceptor distances, increase of overlap integral favored by enhancement of the resonant conditions induced by pressure, and increase of the refractive index. All these magnitudes can be modified at the same time by applying an external pressure at variance with variations of Tm3+ concentration, mainly providing variation of RDA. The present analysis indicates that high-pressure spectroscopy is crucial to unveil the quadrupole-quadrupole interaction as responsible for CR in LiNbO3:Tm3+.

Acknowledgements

This work has been partially supported by Comunidad de Madrid under project MICROSERES-CM (S2009/TIC-1476) and Ministerio de Ciencia e Innovación under project SONAMFIBIOS (MAT2012-34919). We also acknowledge financial support from the Spanish Ministerio de Economia y Competitividad (Project No. MAT2012-38664-C02-1) and MALTA INGENIO-CONSOLIDER 2010(Project Ref. CDS2007-0045). Dr. M. Quintanilla would like to thank Fundación Ramón Areces for financially supporting her through their granting program for Life and Matter Sciences, and J.A. Barreda-Argüeso for a technical grant from the PN de Investigación Científica, Desarrollo e Innovación Tecnológica (Ref. PTA2011-5461-I).

References and links

1. L. Johnson and A. Ballman, “Coherent Emission from Rare Earth Ions in Electro-optic Crystals,” J. Appl. Phys. 40(1), 297–302 (1969). [CrossRef]  

2. J. P. deSandro, J. K. Jones, D. P. Shepherd, M. Hempstead, J. Wang, and A. C. Tropper, “Non-photorefractive CW Tm-indiffused Ti:LiNbO3 waveguide laser operating at room temperature,” IEEE Photon. Technol. Lett. 8(2), 209–211 (1996). [CrossRef]  

3. R. C. Stoneman and L. Esterowitz, “Efficient 1.94-mm-M Tm:YALO Laser,” IEEE J. Sel. Top. Quantum Electron. 1(1), 78–81 (1995). [CrossRef]  

4. V. Sudesh and J. A. Piper, “Spectroscopy, modeling, and laser operation of thulium-doped crystals at 2.3 mm,” IEEE J. Quantum Electron. 36(7), 879–884 (2000). [CrossRef]  

5. T. Kasamatsu, Y. Yano, and T. Ono, “Laser-diode-pumped highly efficient gain-shifted thulium-doped fiber amplifier operating in the 1480-1510-nm band,” IEEE Photon. Technol. Lett. 13(5), 433–435 (2001). [CrossRef]  

6. W. Risk, T. Gosnell, and A. Nurmikko, Compact blue-green lasers (Cambridge University Press, Cambridge, 2003).

7. T. Riedener, H. U. Gudel, G. C. Valley, and R. A. McFarlane, “Infrared to Visible Up-Conversion in Cs3Yb2Cl9-Tm3+,” J. Lumin. 63(5-6), 327–337 (1995). [CrossRef]  

8. E. Saglamyurek, N. Sinclair, J. Jin, J. A. Slater, D. Oblak, F. Bussières, M. George, R. Ricken, W. Sohler, and W. Tittel, “Broadband waveguide quantum memory for entangled photons,” Nature 469(7331), 512–515 (2011). [CrossRef]   [PubMed]  

9. E. Sinclair, E. Saglamyurek, M. George, R. Ricken, C. La Mela, W. Sohler, and W. Tittel, “Spectroscopic investigations of a Ti:Tm:LiNbO3 waveguide for photon-echo quantum memory,” J. Lumin. 130(9), 1586–1593 (2010). [CrossRef]  

10. E. Cantelar, J. A. Sanz-Garcia, G. Lifante, F. Cusso, and P. L. Pernas, “Single polarized Tm3+ laser in Zn-diffused LiNbO3 channel waveguides,” Appl. Phys. Lett. 86(16), 161119 (2005). [CrossRef]  

11. R. A. Hayward, W. A. Clarkson, P. W. Turner, J. Nilsson, A. B. Grudinin, and D. C. Hanna, “Efficient cladding-pumped Tm-doped silica fibre laser with high power singlemode output at 2 mm,” Electron. Lett. 36(8), 711–712 (2000). [CrossRef]  

12. C. Li, J. Song, D. Shen, N. S. Kim, K. Ueda, Y. Huo, S. He, and Y. Cao, “Diode-pumped high-efficiency Tm:YAG lasers,” Opt. Express 4(1), 12–18 (1999). [CrossRef]   [PubMed]  

13. G. Dieke, Spectra and energy levels of rare-earth ions in crystals (Wiley, New York, 1968).

14. M. Quintanilla, E. Cantelar, J. A. Sanz-Garcia, G. Lifante, G. A. Torchia, and F. Cusso, “Infrared energy transfer in Tm3+: LiNbO3,” J. Lumin. 128(5-6), 927–930 (2008). [CrossRef]  

15. M. Quintanilla, E. Cantelar, J. A. Sanz-García, and F. Cusso, “Growth and optical characterization of Tm3+-doped LiNbO3,” Opt. Mater. 30(7), 1098–1102 (2008). [CrossRef]  

16. K. Syassen, “Ruby under pressure,” High Press. Res. 28(2), 75–126 (2008). [CrossRef]  

17. R. A. Forman, G. J. Piermarini, J. D. Barnett, and S. Block, “Pressure Measurement Made by the Utilization of Ruby Sharp-Line Luminescence,” Science 176(4032), 284–285 (1972). [CrossRef]   [PubMed]  

18. J. A. Barreda-Argüeso and F. Rodriguez, “Microscopio para la caracterización espectroscópica de una muestra,” (Patent PCT/ES2014/000049).

19. A. Brenier, C. Pedrini, B. Moine, J. L. Adam, and C. Pledel, “Fluorescence Mechanisms in Tm3+ Singly Doped and Tm3+, Ho3+ Doubly Doped Indium-Based Fluoride Glasses,” Phys. Rev. B Condens. Matter 41(8), 5364–5371 (1990). [CrossRef]   [PubMed]  

20. Y. S. Han, J. H. Song, and J. Heo, “Analysis of cross relaxation between Tm3+ ions in PbO-Bi2O3-Ga2O3-GeO2 glass,” J. Appl. Phys. 94, 2817–2820 (2003). [CrossRef]  

21. B. M. Walsh, N. P. Barnes, D. J. Reichle, and S. Jiang, “Optical properties of Tm3+ ions in alkali germanate glass,” J. Non-Cryst. Solids 352(50-51), 5344–5352 (2006). [CrossRef]  

22. L. Núñez and F. Cusso, “Polarized Absorption and Energy-Levels of LiNbO3:Tm and LiNbO3(MgO)Tm,” J. Phys. Condens. Matter 5(30), 5301–5312 (1993). [CrossRef]  

23. G. De, W. P. Qin, J. S. Zhang, J. S. Zhang, Y. Wang, C. Y. Cao, and Y. Cui, “Infrared-to-ultraviolet up-conversion luminescence of YF3:Yb3+,Tm3+ microsheets,” J. Lumin. 122–123, 128–130 (2007).

24. G. Wang, W. Qin, L. Wang, G. Wei, P. Zhu, and R. Kim, “Intense ultraviolet upconversion luminescence from hexagonal NaYF4:Yb3+/Tm3+ microcrystals,” Opt. Express 16(16), 11907–11914 (2008). [CrossRef]   [PubMed]  

25. R. A. Hewes and J. F. Sarver, “Infrared excitation processes for the visible luminescence of Er3+, Ho3+, and Tm3+ in Yb3+-sensitized rare-earth trifluorides,” Phys. Rev. B Condens. Matter 182, 427–436 (1969).

26. A. Jayaraman and A. A. Ballman, “Effect of pressure on the Raman modes in LiNbO3 and LiTaO3,” J. Appl. Phys. 60(3), 1208–1210 (1986). [CrossRef]  

27. Y. K. Lin, Y. D. Li, Y. W. Xu, G. X. Lan, and H. F. Wang, “Raman-scattering study on pressure amorphization of LiNbO3 crystal,” J. Appl. Phys. 77(7), 3584–3585 (1995). [CrossRef]  

28. L. Riseberg and H. Moos, “Multiphonon Ornit-Lattice Relaxation of Excited States of Rare-Earth Ions in Crystals,” Phys. Rev. 174(2), 429–438 (1968). [CrossRef]  

29. T. Miyakawa and D. L. Dexter, “Phonon sidebands, multiphonon relaxation of excited states, and phonon-assisted energy transfer between ions in solids,” Phys. Rev. B Condens. Matter 1(7), 2961–2969 (1970). [CrossRef]  

30. M. Schuurmans and J. van Dijk, “On Radiative and Non-radiative Decay Times in the Weak Coupling Limit,” Physica B 123(2), 131–155 (1984). [CrossRef]  

31. M. Quintanilla, “Caracterización espectroscópica y aplicaciones de la conversión infrarrojo-visible en LiNbO3 y YF3 activados con iones Tm3+ y Er3+,” PhD Thesis (Universidad Autónoma de Madrid, Madrid, 2010).

32. L. Nunez, G. Lifante, and F. Cusso, “Polarization effects on the line-strength calculations of Er3+-doped LiNbO3,” Appl. Phys. B 62, 485–491 (1996). [CrossRef]  

33. V. Caciuc, A. V. Postnikov, and G. Borstel, “Ab initio structure and zone-center phonons in LiNbO3,” Phys. Rev. B Condens. Matter 61(13), 8806–8813 (2000). [CrossRef]  

34. T. Förster, “Zwischenmolekulare energiewanderung und fluoreszenz,” Ann. Phys. 437(1-2), 55–75 (1948). [CrossRef]  

35. D. L. Dexter, “A theory of sensitized luminescence in solids,” J. Chem. Phys. 21(5), 836–850 (1953). [CrossRef]  

36. R. J. Tonucci, S. M. Jacobsen, and W. M. Yen, “Energy-Transfer Processes in the T-1(2G) and T-3(2G) Excited States of Ni2+:MgO,” Phys. Rev. B Condens. Matter 43(10), 7377–7385 (1991). [CrossRef]   [PubMed]  

37. A. M. Tkachuk, I. K. Razumova, M. F. Joubert, R. Moncorge, D. I. Mironov, and A. A. Nikitichev, “Self-quenching of luminescence in YLF: Tm3+ crystals: I. Microscopic parameters and rates of energy transfer,” Opt. Spectrosc. 85, 885–892 (1998).

38. A. M. Tkachuk, I. K. Razumova, E. Y. Perlin, M. F. Joubert, and R. Moncorge, “Luminescence self-quenching in Tm3+: YLF crystals: II. The luminescence decay and macrorates of energy transfer,” Opt. Spectrosc. 90(1), 78–88 (2001). [CrossRef]  

39. D. F. de Sousa, R. Lebullenger, A. C. Hernandes, and L. A. O. Nunes, “Evidence of higher-order mechanisms than dipole-dipole interaction in Tm3+ → Tm3+ energy transfer in fluoroindogallate glasses,” Phys. Rev. B Condens. Matter 65(9), 094204 (2002). [CrossRef]  

40. D. F. de Sousa and L. A. O. Nunes, “Microscopic and macroscopic parameters of energy transfer between Tm3+ ions in fluoroindogallate glasses,” Phys. Rev. B Condens. Matter 66(2), 024207 (2002). [CrossRef]  

41. J. A. H. Da Jornada, S. Block, F. A. Mauer, and G. J. Piermarini, “Phase-Transition and Compression of LiNbO3 under static high-pressure,” J. Appl. Phys. 57(3), 842–844 (1985). [CrossRef]  

42. M. Hobden and J. Warner, “The temperature dependence of the refractive indices of pure lithium niobate,” Phys. Lett. 22(3), 243–244 (1966). [CrossRef]  

43. T. Kushida, “Energy-transfer and cooperative optical transitions in rare-earth doped inorganic materials.1. Transition probability calculation,” J. Phys. Soc. Jpn. 34(5), 1318–1326 (1973). [CrossRef]  

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 (7)

Fig. 1
Fig. 1 (A) Optical absorption and (B) up-conversion photoluminescence via excitation at 795 nm of LiNbO3:Tm3+ at ambient conditions. On the right a partial scheme of Tm3+ energy states and the selected excitation scheme is added.
Fig. 2
Fig. 2 (A) Pressure dependence of the emission spectrum corresponding to 3H43H6 and 1G43H5 transitions after 476 nm excitation. (B) Decomposition of the emission spectrum at 2.2 GPa into four Lorentzian peaks. The inset shows the variation of the emission peaks with pressure.
Fig. 3
Fig. 3 (A) Variation of the emission spectra of LiNbO3:Tm3+ (2.5 mol%) in the visible and infrared ranges upon cw excitation at 476.5 nm. The infrared 3H43H6 emission peak at 12579 cm−1 (795.0 nm) shown in Fig. 2(a) is missed here because it appears at the spectral limit of the visible spectrometer. (B) Magnification of the variation with pressure of the 3F43H6 and 1G43F3 emission bands, and (C) pressure shifts of levels involved in the CR (3H63F4: 3H43F4).
Fig. 4
Fig. 4 Variation of the relative intensity of the main emission peaks of LiNbO3:Tm3+ with pressure. The relative variation for each peak is compared with the variation of the 3H43H6 peak taken as reference.
Fig. 5
Fig. 5 (A) Intensity decay curves corresponding to the 3H43H6 emission (795 nm) after pulsed excitation into the1G4 state (475 nm) in LiNbO3:Tm3+ (2.5 mol%) at different pressures. The decays can be fitted to single exponential functions (blue solid lines). (B) Pressure dependence of the associated lifetime calculated by fitting the experimental data I(t) to a single-exponential function. The grey dotted line is a linear fit to the data, added to guide the eye.
Fig. 6
Fig. 6 Dependence of the measured transfer probability on the hydrostatic pressure. The experimental WET data (blue points), have been obtained from the measured lifetime using Eq. (1), taking τW0 = 230 μs. The blue dashed-dotted line has been drawn to guide the eye. The figure includes the probabilities, calculated from Eqs. (3) and (5), for the three main orders of the interaction (d-d, d-q and q-q). (A) The calculated probabilities consider only variations in RDA due to pressure-induced volume compression by means of the LiNbO3 equation of state. (B) Calculated energy-transfer probabilities due to both variations in RDA and refractive index of LiNbO3.
Fig. 7
Fig. 7 (A) Spectral overlap of the donor emission and acceptor absorption in the considered cross-relaxation process. (B) Experimental WET (blue points) and calculated (orange dotted line) predictions for quadrupole–quadrupole interaction considering all possible contributions: RDA variations, refractive index changes and spectral overlap modification. The grey dashed lines have been calculated considering a ±0.1 cm−1/GPa error estimation in the shift coefficients of the peaks in order to illustrate the uncertainty of the calculation.

Tables (1)

Tables Icon

Table 1 Peak position at ambient pressure, ωi(P0), and linear pressure shifts, (∂ωi/∂P), of the emission peaks associated to all the measured transitions. The most intense ones are highlighted in bold type.

Equations (6)

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

1 τ exp = 1 τ rad + 1 τ NR + W ET = 1 τ W 0 + W ET
W NR = β el exp[ α( ΔE2ω ) ]
W ET = C DA (S) R DA S
C DA (6) = 3 4 c 4 Q A 4π n 4 τ 0 L D ( E ) L A (E) E 4 dE; C DA (8) = 135α 6 c 6 Q A 4 πn 4 τ 0 g D g A * g A g D * L D ( E ) L A (E) E 6 dE; C DA (10) = 225ε 8 c 8 Q A 4π n 4 τ 0 L D ( E ) L A (E) E 8 dE
V V 0 = ( R R 0 ) 3 = ( 1+ B 0 ' ( P B 0 ) ) 1 B 0 '
ΔV V = 6n ( n 2 1 )( n 2 +2 ) Δn
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.