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

Excited-state absorption in thulium-doped materials in the near-infrared

Open Access Open Access

Abstract

Excited-state absorption (ESA) is a key process for upconversion pumping schemes of thulium (Tm3+) doped laser materials. We have systematically studied two ESA transitions in the near-infrared spectral range, namely 3F43F2,3 (at ∼1 µm) and 3F43H4 (at ∼1.5 µm), in various Tm3+-doped fluoride (ZBLAN glass, cubic KY3F10 and CaF2, tetragonal LiYF4 and LiLuF4, monoclinic BaY2F8 crystals) and oxide (cubic Y3Al5O12, orthorhombic YAlO3 crystals) laser materials, using a pump-probe method with a polarized light. An approach to calculate the constants of energy-transfer upconversion (ETU) is also presented.

© 2022 Optica Publishing Group under the terms of the Optica Open Access Publishing Agreement

1. Introduction

Excited-state absorption (ESA) is a process of excitation of a single atom, ion or molecule from a lower-lying excited-state |1 > to a higher-lying one |2 > with the absorption of a photon [1]. ESA can occur only if the system has been already excited to the |1 > state from the ground-state |0 > by ground-state absorption (GSA), Fig. 1. Therefore, the excitation to the |2 > state via ESA involves at least two pump photons. The energy differences between the (|0 > and |1>) and (|1 > and |2>) states can be either similar (resonant) or different. In the latter case, two pump beams with different wavelengths are required to observe ESA. ESA is commonly observed in molecules and materials doped with transition metal (TM) ions [24]. It is less common for rare-earth (RE3+) doped materials because of their narrow bandwidth transitions. Among RE3+ ions, a larger variety of ESA processes occurs for those exhibiting rich (ladder-like) energy-level schemes [57].

 figure: Fig. 1.

Fig. 1. Two-photon processes: (a) excited-state absorption, (b) energy-transfer upconversion and (c) two-photon absorption. |0>, ground-state, |1 > and |2>, excited-states, |1'>, virtual state.

Download Full Size | PDF

ESA should be distinguished from other two-ion or two-photon processes, such as energy-transfer upconversion (ETU) and two-photon absorption (TPA). For ETU, two neighboring centers are both excited to the |1 > state and transfer non-radiatively the excitation energy, so that one ion is further excited to the higher-lying |2 > state and another one returns to the ground-state |0> [8], Fig. 1. The probability of ETU depends on the concentration of centers and their spatial distribution [9,10]. TPA is a process of simultaneous absorption of two pump photons with the same frequency leading to the population of an excited state |1 > through a virtual state |1'>, Fig. 1. It is a third-order process, and its probability depends quadratically on the light intensity.

Frequently, ESA is used to obtain emission from higher-lying excited-states with a frequency νe being higher than the pump photon frequency νp (an anti-Stokes process). ESA can also assist in obtaining Stokes emissions from such higher lying states. In this way, ESA can be a key process for upconversion pumping of RE3+-doped laser materials resulting in lasing in the infrared [11,12] or visible [13,14] spectral ranges. Concerning the latter, upconversion pumping is known for several rare-earth ions, such as Er3+, Pr3+, Tm3+ [15].

If the transitions |0> → |1 > and |1> → |2 > are not resonant in energy, the higher-lying excited-state can be populated either by using two different pump beams with photon frequencies νp1 and νp2, or via the photon avalanche (PA) effect [16]. PA was first evidenced by Chivian et al. for Pr3+ ions [17]. Furthermore, it was used to demonstrate upconversion pumping of laser materials doped with Pr3+ [18], Nd3+ [19] and Tm3+ [20] ions. The typical scheme of PA is illustrated in Fig. 2. The first step is a non-resonant (phonon-assisted) GSA to the |1 > state followed by a resonant ESA, |1> → |2 > . The second step relies on the cross-relaxation (CR) process between two adjacent ions, one being in the higher-lying excited-state |2 > and one - in the ground-state |0 > . Because of the non-radiative energy transfer, the first one experiences a de-excitation to the intermediate state |1 > and the second one is promoted to the same state. As a consequence, the population of the |1 > state increases thus feeding the ESA process. In this way, the higher lying |2 > state is efficiently populated even if the first GSA transition is non-resonant.

 figure: Fig. 2.

Fig. 2. Photon avalanche effect scheme: GSA and ESA – ground- and excited-state absorption, respectively, CR – cross-relaxation, -ph and + ph – absorption / generation of phonons.

Download Full Size | PDF

Among the rare-earth ions, thulium (Tm3+) is attracting a lot of attention for coherent light generation in the short-wavelength-infrared (SWIR) spectral range. Tm3+ has an electronic configuration of [Xe]4f12 and its simplified scheme of energy levels is shown in Fig. 3(a). The most widely exploited Tm3+ laser transition originates from the metastable 3F4 state (3F43H6, at ∼2 µm) [21] and less common ones – from the higher-lying 3H4 state (3H43F4, at ∼1.5 µm and 3H43H5, at ∼2.3 µm [22]).

 figure: Fig. 3.

Fig. 3. Thulium ions: (a) simplified energy-level scheme showing laser transitions in the near- and mid-infrared; (b-c) pumping schemes of Tm3+ ions: (b) conventional pumping; (c) upconversion pumping, GSA and ESA – ground- and excited-state absorption, CR - cross-relaxation, NR – multiphonon non-radiative relaxation.

Download Full Size | PDF

To achieve laser emission in all these channels, Tm3+ ions are typically excited to the 3H4 state owing to the intense 3H63H4 GSA transition spectrally overlapping with the emission of commercial AlGaAs laser diodes and Ti:Sapphire lasers. The CR process, 3H4 + 3H63F4 + 3F4, Fig. 3(b), is very efficient in many Tm3+-doped materials even at moderate doping levels, resulting in efficient population of the 3F4 metastable state with a pump quantum efficiency approaching 2 (a two-for-one pump process) [23] and giving rise to high laser efficiencies particularly at ∼2 µm well exceeding the Stokes limit [24].

Owing to the complex ladder-like structure of Tm3+ energy levels, the presence of efficient CR process and a long-living metastable level, the upconversion pumping of Tm3+-doped materials relying on the PA effect for lasing at ∼2.3 µm was also realized recently [20,25], Fig. 3(c). It is based on the non-resonant GSA 3H63H5 followed by multiphonon NR relaxation to the metastable level 3F4, followed by a resonant ESA 3F43F2,3 and efficient CR. Altogether, this leads to efficient population of both the 3F4 and 3H4 states.

As explained above, excited-state absorption is a key effect for realization of upconversion pumping of Tm lasers emitting in the mid-infrared, at ∼2.3 µm. Such lasers find applications in sensing of atmospheric species such as HF, CO, CH4 or H2CO, non-invasive glucose (C6H12O6) blood measurements, or frequency conversion further in the mid-infrared. Unfortunately, to date, the information about the near-infrared ESA properties of commonly used Tm3+-doped laser crystals is relatively scarce.

In the present work, we aimed to perform a systematic study of excited-state absorption of Tm3+ ions in the near-infrared for a series of laser host materials (fluorides and oxides).

2. Methodology

2.1 Excited-state absorption by thulium ions

Figure 4 summarized the excited-state absorption transitions of Tm3+ ions exploited so far. The corresponding ground-state absorption transitions are shown as well. Note that we consider the case of resonant energy differences between the (|0 > and |1>) and (|1 > and |2>) states, Fig. 1(a). The transition 3F41D2 falls in the blue spectral range (∼0.45 µm) and is resonant with the 3H61G4 GSA channel. The transitions 3H41D2 and 3F41G4 both in the red are nearly resonant with the 3H63F2,3 GSA channel; they are crucial for demonstrating upconversion pumping of blue thulium lasers [2628]. The ESA properties of various Tm3+ doped materials in the red spectral range have been extensively studied in the past [5,29].

 figure: Fig. 4.

Fig. 4. Summary of excited-state absorption transitions of thulium ions exploited so far (bold arrows – this work), the corresponding ground-state absorption is also shown.

Download Full Size | PDF

In the near-IR, the ESA transitions 3F43F2,3 and 3H41G4 at ∼1 µm are close in energy with the 3H63H5 GSA channel. The 3F43H4 ESA transition at ∼1.5 µm is close to the 3H63H5 GSA channel. In the present work, we focus on two near-infrared ESA transitions, 3F43F2,3 and 3F43H4, which are relevant for UC pumping of mid-infrared Tm lasers. So far, these ESA channels have been studied only for Tm:Y3Al5O12 [30]. Some incomplete ESA data were also reported for Tm:KY3F10 and Tm:LiYF4 [31].

2.2 Experimental setup

The measurements of the ESA spectra of Tm3+-doped materials corresponding to the 3F43F2,3 and 3F43H4 transitions were carried out using a pump-probe method [1,32]. Its principle is the following: the first intense laser beam (the pump) excites Tm3+ ions into the metastable 3F4 state. The second weak beam (the probe) propagates through the pumped area of the sample revealing the additional absorption caused by ESA.

The scheme of the pump-probe set-up is shown in Fig. 5. The sample was prepared as a relatively thin (a few mm thick) plane-parallel plate with both faces polished to laser quality to avoid scattering of the pump. It was placed between two pinholes (diameter: 500 µm) placed close to the sample faces. This provided a good spatial overlap between the pump and the probe beams, as well as helped to limit the spot size of the probe in the sample.

 figure: Fig. 5.

Fig. 5. (a) Scheme of the pump-probe setup used for ESA measurements: AOM, acousto-optic modulator, M1 and M2, folding mirrors, L1, focusing lens, L2 – L4, wide-aperture lenses, νpump and νprobe, pump and probe modulation frequencies, respectively, P, polarizer, F, filter; (b) Energy-level scheme of Tm3+ ions illustrating the principle of the pump-probe method.

Download Full Size | PDF

As a source of pump radiation, we used a continuous-wave Ti:Sapphire laser (Coherent 890) delivering up to 1.5 W at ∼0.8 µm in the fundamental transverse mode (M2 ≈ 1). The pump wavelength was tuned to match precisely the maximum of the 3H63H4 Tm3+ absorption band of the studied material. The pump beam was modulated at low frequency (νpump ≈ 10 Hz) by an acousto-optic modulator (Isomet, model 1205C-2) and it was focused into the sample using a lens L1 (f = 150 mm) resulting in a spot size (diameter) of ∼300 µm. The single-pass pump absorption was >50%. This resulted in a relatively uniform distribution of inversion through the sample.

A 100 W halogen-tungsten white-light lamp (Oriel, model 66184) with a broad emission spectrum close to that of a black body was used as a probe source. Its output was modulated at a higher frequency (νprobe ≈ 1 kHz) by a mechanical chopper. The probe beam was also focused into the sample using a pair of wide-aperture lenses L2 and L3 (f = 50 mm and 100 mm, respectively) and after passing the sample, it was reimaged on the input slit of the monochromator using the lens L4 (f = 100 mm).

For measuring the spectra, a 0.64 m monochromator (Jobin-Yvon, HRS2) and an InGaAs detector with two lock-in amplifiers (SR810 DSP, Stanford Research Systems) were used. The first one was locked with the pump modulation frequency νpump and measured the average intensity transmitted through the sample I and the second one was locked with the probe frequency νprobe to measure the pump-induced variation ΔI of the abovementioned transmission. For the developed setup, the transmission variation ΔI/I was about ∼10−4…10−3 at the wavelengths of the ESA peaks.

A Glan-Taylor polarizer was placed before the input slit of the monochromator for polarization-resolved measurements. A long-pass filter (LP950, Spectrogon) was also used to filter out the residual pump radiation.

The wavelength calibration was performed with a mercury vapor lamp (Schwabe, Sp 60/U-V). The measurements were performed in the spectral range of 950-1575 nm with a resolution (spectral bandwidth, SBW) of 1-2 nm, depending on the sample. All the studies were performed at room temperature (RT, 20 °C).

2.3 Materials

Table 1 presents the list of the Tm3+-doped fluoride (ZBLAN glass, KY3F10, CaF2, LiYF4, LiLuF4 and BaY2F8 crystals) and oxide (Y3Al5O12 and YAlO3 crystals) laser materials that have been studied.

Tables Icon

Table 1. Summary of the Studied Thulium-Doped Materials

To ensure that no ESA from the 3H4 pump level (e.g., the 3H41G4 transition at ∼1 µm, cf. Figure 4) can contribute to the measured spectra, we used highly Tm3+-doped materials (several at.% Tm, Table 1) providing efficient self-quenching of the 3H4 lifetime by cross-relaxation. Under these conditions, a favorable ratio of the 3F4 to 3H4 lifetimes (τlum(3F4) of about few ms and τlum(3F4) of about few tens of µs) was observed.

For the optically isotropic glass and cubic crystals (KY3F10, CaF2 and Y3Al5O12), the studies were performed using unpolarized light. For the optically uniaxial crystals (LiYF4 and LiLuF4, the optical axis is parallel to the c-axis), the measurements were performed using a-cut samples for the two principal light polarizations, π (E || c) and σ (Ec). For the optically biaxial crystals (BaY2F8 and YAlO3), several samples with different cuts were used giving access to three principal light polarizations. For monoclinic BaY2F8, they were denoted as E || X, Y and Z (X, Y and Z are the optical indicatrix axes; their assignment is given in [33]). For orthorhombic YAlO3, the optical indicatrix frame coincides with the crystallographic one, so that the polarizations were denoted simply as E || a, b and c. Here, the crystallographic axes are selected according to the standard crystallographic setting Pnma [34].

2.4 Methodology

The relative variation of the transmitted probe signal ΔI/I is given by [32]:

$$\frac{{\Delta I}}{I}(\lambda ) = A{N^\ast }L({\sigma _{GSA}}(\lambda ) + {\sigma _{SE}}(\lambda ) - {\sigma _{ESA}}(\lambda )),$$
where A is a calibration factor, L is the optical length of the sample, N* is the population of the probed excited-state, the 3F4 state of Tm3+, in our case (these three parameters are independent of the wavelength and they are determined by the geometry of the pump-probe experiment), σGSA and σESA are the ground- and excited-state absorption cross-sections, respectively, σSE is the stimulated-emission (SE) cross-section. The σGSA and σSE values are found independently and the corresponding spectral bands can be used for calibrating the set-up (i.e., determining the constant value of AN*L for each particular sample). In this way, the spatial and longitudinal distributions of the population of the metastable excited-state |1 > do not affect the measured ESA spectra. Equation (1) represents the first-order development of the transmission of the probe light and its is valid in the case of relatively weak ΔI/I variations [32].

There are two Tm3+ emissions falling in the studied spectral range, namely 3H43F4 (at ∼1.5 µm) and 3F43H6 (at ∼2 µm). The former emission originates from the 3H4 state which is different from the probed one (3F4). The spectral overlap of the ESA spectra extending from ∼1 to 1.6 µm with the second emission is very weak. Thus, for simplicity, we assume σSE ≈ 0 in Eq. (1).

For the 3F43H4 transition (at ∼1.5 µm), the ESA cross-sections can be independently calculated using the reciprocity method [35] from the σSE spectra of the 3H43F4 transition:

$${\sigma _{ESA}}(\lambda ) = {\sigma _{SE}}(\lambda )\frac{{{Z_2}}}{{{Z_1}}}\exp \left[ {\frac{{(hc/\lambda ) - {E_{ZPL}}}}{{kT}}} \right],$$
where h the Planck constant, c is the speed of light, λ is the light wavelength (hc/λ is the photon energy), k is the Boltzmann constant, T the sample temperature (RT), EZPL is the energy of the zero-phonon-line (ZPL) transition between the lowest Stark sub-levels of both multiplets, and Z1(2) are the partition functions of the lower and upper multiplets, respectively:
$${Z_m} = \sum\nolimits_k {g_k^m} {e^{ - \frac{{E_k^m}}{{kT}}}},$$
where each Stark sub-level has a number k, an energy Emk measured from the lowest sub-level of each multiplet and a degeneracy gmk.

The ESA spectra for the 3F43F2,3 transition cannot be calculated by the reciprocity method because the 3F2,3 states are depopulated by efficient multiphonon NR relaxation (the energy-gap to the lower-lying 3H4 state is ∼1800cm-1) and does not exhibit detectable luminescence.

2.5 Methodology: the case study of Tm:LiYF4

To illustrate the analysis procedure of measuring and interpreting the Tm3+ ESA spectra, we have selected the Tm:LiYF4 crystal. So far, it is the most widely used material for laser operation of Tm3+ ions in the mid-infrared. Figure 6 shows the raw spectrum, σGSAσESA, proportional to the ΔI/I ratio, the independently determined σGSA spectra for the 3H63H5 and 3H63F4 transitions (obtained from classical absorption studies) and the σSE spectra for the 3H43H5 and 3F43H6 transitions (obtained from emission spectra). The same figure shows the ESA spectra determined using Eq. (1). Following state-of-the-art studies [1], they are plotted in the range of negative cross-sections, i.e., as (–σESA).

 figure: Fig. 6.

Fig. 6. Evaluation of the ESA cross-section spectra, σESA, for the Tm:LiYF4 crystal in the near-infrared: violet – calibrated raw spectrum ∼ΔI/I, blackσGSA spectra, greenσSE spectra, red - σESA spectra. The light polarization is π.

Download Full Size | PDF

To confirm our assignment of the measured ESA spectral bands, we have calculated the full set of wavelengths corresponding to electronic transitions between the Stark sub-levels of the involved Tm3+ multiplets (3F4 and 3H4 + 3F2,3). For this, the data on the crystal-field splitting for Tm3+ ions in LiYF4 were used [36], Table 2. Here, we use empirical notations for Stark sub-levels (3F4 = Yi, 3H4 = Wj, 3F3 = Vk, 3F2 = Um, where the subscripts i, j, etc., number the sub-levels starting from 1) proposed by Lupei et al. [37], as well as indicate the irreducible representations (Г1, Г2 or Г3,4) for each sub-level.

Tables Icon

Table 2. Crystal-Field Splitting of Selected Tm3+ Multiplets in LiYF4

The polarization selection rules for electric-dipole (ED) and magnetic-dipole (MD) transitions for ions in S4 sites are listed in Table 3. The MD transitions are those following the selection rules ΔJ = |J – J'| = 0, ±1, except of 0 ↔ 0’. In our case, the ESA transitions 3F43F3 and 3F43H4 can have a MD contribution. Indeed, the Judd-Ofelt calculations for Tm3+ ions in LiYF4 predict the following ED and MD probabilities of spontaneous radiative transitions: AED = 51.35 s-1 and AMD = 18.95 s-1 (for 3H43F4) and AED = 44.09 s-1 and AMD = 35.49 s-1 (for 3F33F4) [38]. Thus, the MD contribution to the ESA probabilities is expected to be non-negligible, especially for the 3F43F3 transition. The extra MD lines should correspond to transitions between the Stark sub-levels having the same irreducible representations, i.e., Г1 → Г1 and Г2 → Г2, only for π-polarization, cf. Table 3.

Tables Icon

Table 3. Polarization Selection Rules for Electric- and Magnetic-Dipole Transitions in S4 Sites

The results on the assignment of the measured ESA spectra of Tm3+ ions to ED and MD allowed electronic transitions are shown in Fig. 7. Several conclusions can be drawn from this figure. First, the observed ESA lines are well assigned only to transitions originating from the 3F4 state and terminating at the 3H4 and 3F2,3 ones (i.e., extra lines due to possible ESA from the 3H4 state are not observed). Second, the polarization selection rules well explain the polarization anisotropy of ESA spectra. Third, the prominent ESA peaks can be in most cases assigned to single electronic transitions which determine their narrow-linewidth nature. Moreover, the ESA transitions terminating at the closely located 3F3 and 3F2 states are well resolved.

 figure: Fig. 7.

Fig. 7. Interpretation of the polarized ESA cross-section, σESA, spectra for Tm3+ ions in the LiYF4 crystal in the near-infrared: (a,c) the 3F43F2,3 transitions, (b,d) the 3F43H4 transition, light polarizations are (a,b) π and (c,d) σ. Curves – measured spectra, in (c,d), spectra calculated using the reciprocity method (RM) are shown for comparison, vertical dashes – ED electronic transitions of Tm3+ ions according to the polarization selection rules.

Download Full Size | PDF

Figure 7(b,d) also contains the ESA cross-section spectra for the 3F43H4 transition calculated by means of the reciprocity method. They are in good agreement with the measured ones confirming the correctness of the pump-probe-based approach.

3. Excited-state absorption spectra

3.1 Fluoride materials

The ESA cross-section spectra corresponding to the 3F43F2,3 and 3F43H4 transitions for Tm3+ ions in fluoride materials (ZBLAN glass, cubic KY3F10 and CaF2, tetragonal LiYF4 and LiLuF4 and monoclinic BaY2F8 crystals) are shown in Fig. 8. Let us briefly discuss them.

 figure: Fig. 8.

Fig. 8. Polarized (where applicable) ESA cross-section, σESA, spectra for Tm3+ ions in fluoride materials: (a, b) isotropic hosts: ZBLAN glass, cubic KY3F10 and CaF2 crystals; (c-f) uniaxial crystals: (c,d) tetragonal LiYF4 and (e,f) LiLuF4 and (g,h) biaxial crystal: monoclinic BaY2F8. Transitions: (a,c,e,g) 3F43F2,3 and (b,d,f,h) 3F43H4. The spectral resolution is indicated on the graphs (it is the same in left and right panels).

Download Full Size | PDF

Due to its amorphous nature, the Tm:ZBLAN glass shows smooth and broad bands at the expense of low peak ESA cross-sections. For the 3F43F2,3 ESA transition which is of interest for UC pumping at ∼1 µm, σESA is 0.23×10−20 cm2 at 1056 nm and the corresponding absorption bandwidth (full width at half maximum, FWHM) ΔλESA is 39 nm. As Tm:ZBLAN glasses are used in the form of fibers, relatively high absorption via UC pumping is easily achievable. A ∼2.3 µm Tm:ZBLAN fiber lasers with UC pumping at ∼1 µm was recently reported [25].

Among the studied Tm3+-doped fluoride crystals, Tm:CaF2 is standing apart: its spectra are similar to those of the glass. Such a “glassy-like” behavior is due to the strong ion clustering even at low doping levels leading to significant inhomogeneous spectral broadening [39]. For the 3F43F2,3 ESA transition, σESA is low, only 0.07×10−20 cm2 at 1064 nm with broad ΔλESA of 38 nm. The origin of the ESA peak at 1089.7 nm is not clear; absorption of residual isolated Tm3+ ions in cubic symmetry sites (Oh) can explain this sharp line.

Other Tm3+-doped fluoride crystals exhibit much narrower ESA peaks owing to electronic transitions of ions located in a single type of sites with reduced inhomogeneous broadening (cf. Table 1). For cubic Tm:KY3F10, σESA reaches 0.91×10−20 cm2 at 1067.7 nm with ΔλESA = 3.4 nm. Another intense ESA line is observed at 1048.7 nm corresponding to a σESA value of 0.73×10−20 cm2 and a ΔλESA of 3.6 nm. This line was recently used for demonstrating UC pumping of a ∼2.3 µm Tm:KY3F10 laser [40].

For optically uniaxial and biaxial fluoride crystals, a noticeable polarization-anisotropy of ESA spectra is observed. For the 3F43F2,3 ESA transition of Tm3+ ions in tetragonal LiYF4 and LiLuF4 crystals, higher ESA cross-sections are measured for σ-polarized light: σESA reaches 0.62×10−20 cm2 at 1042.0 nm and 0.36×10−20 cm2 at 1055.9 nm with ΔλESA = 2.5 nm and 3.2 nm, respectively (Tm:LiYF4, σ-polarization). Both these ESA lines were used for UC pumping of a ∼2.3 µm Tm:LiYF4 laser [20]. The ESA properties of the isostructural LiY/LuF4 crystals are similar, while the ESA cross-sections seem to be slightly higher for the Lu-compound. Finally, for monoclinic Tm:BaY2F8, the preferable light polarization for UC pumping at ∼1 µm is E || Z, as the corresponding σESA reaches 0.27×10−20 cm2 at 1047.2 nm with ΔλESA = 5.5 nm.

Among the studied fluoride materials, Tm:KY3F10 and Tm:LiY/LuF4 crystals offer the most intense ESA lines around ∼1 µm making them attractive for UC pumping.

3.2 Oxide materials

The ESA cross-section spectra corresponding to the 3F43F2,3 and 3F43H4 transitions for Tm3+ ions in oxide crystals (cubic Y3Al5O12 and orthorhombic YAlO3) are shown in Fig. 9.

 figure: Fig. 9.

Fig. 9. Polarized (where applicable) ESA cross-section, σESA, spectra for Tm3+ ions in oxide crystals: (a,b) cubic Y3Al5O12 and (c,d) orthorhombic YAlO3 crystals. Transitions: (a,c) 3F43F2,3 and (b,d) 3F43H4. The spectral resolution is indicated on the graphs.

Download Full Size | PDF

For cubic Tm:Y3Al5O12, the 3F43F2,3 ESA transition, σESA reaches 0.48×10−20 cm2 at 1031.0 nm with a relatively narrow ΔλESA of 1.5 nm.

For orthorhombic Tm:YAlO3, the ESA spectra exhibit a significant polarization anisotropy. The most attractive polarization is E || a for which σESA is 1.55×10−20 cm2 (the highest value among the studied materials) at 1044.2 nm corresponding to ΔλESA of 2.7 nm.

3.3 Towards upconversion pumping at ∼1 µm

Table 4 lists the ESA peaks slightly above 1 µm (the 3F43F2 transition) which are of practical interest for realization of upconversion pumping of ∼2.3 µm Tm lasers. The corresponding pump sources can be commercially available Yb bulk or fiber lasers offering power-scalable output as well as wavelength tunability in this spectral range.

Tables Icon

Table 4. Excited-State Absorption Peaks* of Tm3+ Ions in Fluoride and Oxide Materials Suitable for Upconversion Pumping at ∼1 µm

4. Energy-transfer upconversion

The ESA spectra in the IR spectral range can be also used for evaluating the concentration-independent parameter of one of the energy-transfer upconversion (ETU) processes for Tm3+ ions, i.e., 3F4 + 3F43H4 + 3H6. This process acts against the cross-relaxation, 3H4 + 3H63F4 + 3F4, refilling the higher-lying 3H4 excited-state and depopulating the metastable 3F4 state. It can significantly affect the performance of Tm lasers. In particular, it is detrimental for Tm lasers operating on the 3F43H6 transition (at ∼1.9 µm) [41]: strong ETU in highly Tm3+-doped crystals is responsible for the increased laser threshold and additional heat generation possibly leading to higher risk of thermal fracture. In opposite, this ETU process could be useful for Tm lasers operating on the 3H43H5 transition (at ∼2.3 µm) as it refills the upper laser level at the expense of the metastable intermediate 3F4 state [42]. In this way, pump quantum efficiencies up to 2 could be reached leading to laser slope efficiencies exceeding the Stokes limit. Thus, knowledge of the ETU parameter is of practical importance. Below, we describe the general calculation procedure and apply it to the particular case of Tm:LiYF4.

The ETU rate was evaluated using the hopping (Burshtein) model describing a migration-assisted energy transfer [9]. This model treats the transfer of energy between donors as a random process until the excitation is transferred to an acceptor ion. First, two microparameters CDD and CDA were calculated from the overlap integrals between the SE and absorption (GSA or ESA) cross-section spectra [43]:

$${C_{DD}} = \frac{{3c}}{{8{\pi ^4}{n^2}}}\int {{\sigma _{SE}}(\lambda )} {\sigma _{GSA}}(\lambda )d\lambda ,$$
$${C_{DA}} = \frac{{3c}}{{8{\pi ^4}{n^2}}}\int {{\sigma _{ESA}}(\lambda )} {\sigma _{SE}}(\lambda )d\lambda ,$$
where D and A indicate a donor and an acceptor, respectively, CDD represents a donor-donor process (energy migration) and CDA – a donor-acceptor process (direct energy transfer), Fig. 10, c the speed of light in vacuum, n the mean refractive index of the host, σSE corresponds to the 3F43H6 transition, σGSA and σESA – to the transitions 3H63F4 (GSA) and 3F43H4 (ESA), respectively, λ is the light wavelength. For an anisotropic crystal, the transition cross-sections should be polarization averaged. E.g., for Tm:LiYF4, this means <σ> = (2σσ + σπ)/3, where the subscripts σ and π indicate the polarizations. In our case, the use of the hopping model is validated because CDD >> CDA (i.e., the migration process among Tm3+ ions is more likely than direct energy transfer). Indeed, for Tm:LiYF4, CDD = 3.26×10−39 cm6s-1 and CDA = 0.50×10−43 cm6s-1.

 figure: Fig. 10.

Fig. 10. Energy-transfer upconversion (ETU), 3F4 + 3F43H4 + 3H6, for Tm3+ ions. Two elementary processes are shown: (a) energy-migration and (b) phonon-assisted “direct” ETU.

Download Full Size | PDF

The macroscopic ETU rate in (s-1) was then determined as [44]:

$${W_{ETU}} = \pi {\left( {\frac{{2\pi }}{3}} \right)^{5/2}}\sqrt {{C_{DD}}{C_{DA}}} N_{Tm}^2,$$
where NTm is the doping concentration. There exist several ways to express the dependence of the ETU rate on NTm [23]:
$${W_{ETU}} = {K_{ETU}}{N_{Tm}} = {C_{ETU}}N_{Tm}^2,$$
where KETU and CETU are the concentration-dependent and concentration-independent ETU parameters expressed in (cm3s-1) and (cm6s-1), respectively. When the migration process is very fast because of a high donor concentration, the KETU parameter is expected to be independent on the doping level (KETU = const). However, for most gain materials, the donor concentration typically stays below 20 at.% and the KETU parameter is proportional to the donor concentration, as expressed by Eq. (6). This situation is called migration-limited energy transfer [9].

The considered ETU process is non-resonant and is phonon-assisted. The calculation of the overlap integrals between the SE and absorption (GSA and ESA) spectra should account for the multiphonon sidebands. The shape of the latter can be expressed as [45,46]:

$${\sigma _\textrm{S}} = {\sigma _0}\exp ( - {\alpha _\textrm{S}}\Delta E),$$
$${\sigma _{\textrm{AS}}} = {\sigma _0}\exp ( - {\alpha _{\textrm{AS}}}\Delta E),$$
where S and AS indicate Stokes and anti-Stokes process, respectively, ΔE is the energy mismatch between the vibronic (phonon-assisted) and the purely electronic transition, σ0 is the cross-section at the photon energy of an electronic transition and the constants αS and αAS are determined by vibronic properties of the host material. For Tm:LiYF4, we determined them in the present work to be αS = 7.9 ± 1×10−3 cm-1 and αAS = 12.1 ± 2×10−3 cm-1.

Using Eqs. (4)–(6) and the spectroscopic data for Tm:LiYF4, as shown in Fig. 11(a), we calculated the concentration-independent ETU parameter, CETU = 2.56×10−40 cm6s-1. In the previous studies, the ETU parameter (KETU) was typically reported for given Tm3+ doping concentrations, Fig. 11(b). In this work, the result of our calculation can be expressed as a linear dependence of KETU on NTm with a slope equal to CETU. Let us briefly review the methods used in previous studies to obtain the KETU values. So far, they were obtained from (i) luminescence-decay measurements [47,48], (ii) stationary luminescence-intensity measurements [42], (iii) theoretical calculations [49], and (iv) modeling of Tm laser performance [50,51]. Our results are in good agreement with the previously reported data, especially with the experimental ones from [42].

 figure: Fig. 11.

Fig. 11. Evaluation of the ETU parameter for Tm:LiYF4: (a) polarization-averaged SE (3F43H6), GSA (3H63F4) and ESA (3F43H4) spectra plotted in semi-log scale, solid curves – measured spectra with phonon sidebands calculated using Eq. (7); (b) summary of the ETU parameters KETU reported so far: symbols – literature data, line – this work, calculation using Eqs. (4)–(6).

Download Full Size | PDF

The determined CETU value should be compared with that for the cross-relaxation process, i.e., CCR = 0.25 ± 0.03×10−37 s−1cm6 for Tm:LiYF4 [42]. One can see that the considered ETU process is by two orders of magnitude weaker that the CR one (for each particular Tm3+ doping level).

5. Conclusion

To conclude, we have studied excited-state absorption of Tm3+ ions in the near-infrared spectral range for a series of fluoride and oxide laser crystals using a dedicated pump-probe method. The considered ESA transitions originate from the metastable 3F4 Tm3+ state and terminate at the higher lying 3F2,3 and 3H4 levels. The 3F43F2,3 ESA channel is particularly attractive for upconversion pumping of Tm lasers via a photon avalanche mechanism. This pumping scheme can be used for obtaining laser emission at ∼2.3 µm corresponding to the 3H43H5 transition with commercial, power scalable and wavelength tunable Yb fiber lasers emitting slightly above 1 µm. The ESA line positions, cross-sections and spectral linewidths which are required for the design of upconversion pumped Tm lasers were measured in detail. Among the studied crystals, Tm:KY3F10, Tm:LiY/LuF4 and Tm:YAlO3 are identified as promising candidates for upconversion pumping. Regarding the second studied ESA channel, 3F43H4, we show a route for calculating another key spectroscopic parameter for the development of Tm lasers, namely, the energy-transfer upconversion parameter. The ETU parameters are calculated from the absorption and emission overlap integrals using the hopping model.

Funding

European Science Foundation (NOVAMAT); Région Normandie (NOVAMAT); Agence Nationale de la Recherche (ANR-19-CE08-0028).

Disclosures

The authors declare no conflicts of interest.

Data Availability

Data underlying the results presented in this paper are not publicly available at this time but may be obtained from the authors upon reasonable request.

References

1. J. Koetke and G. Huber, “Infrared excited-state absorption and stimulated-emission cross sections of Er3+-doped crystals,” Appl. Phys. B 61(2), 151–158 (1995). [CrossRef]  

2. N. V. Kuleshov, V. G. Shcherbitsky, V. P. Mikhailov, S. Kück, J. Koetke, K. Petermann, and G. Huber, “Spectroscopy and excited-state absorption of Ni2+-doped MgAl2O4,” J. Lumin. 71(4), 265–268 (1997). [CrossRef]  

3. K. V. Yumashev, I. A. Denisov, N. N. Posnov, N. V. Kuleshov, and R. Moncorge, “Excited state absorption and passive Q-switch performance of Co2+ doped oxide crystals,” J. Alloys Compd. 341(1-2), 366–370 (2002). [CrossRef]  

4. Z. Burshtein, P. Blau, Y. Kalisky, Y. Shimony, and M. R. Kikta, “Excited-state absorption studies of Cr4+ ions in several garnet host crystals,” IEEE J. Quantum Eletron. 34(2), 292–299 (1998). [CrossRef]  

5. N. Garnier, R. Moncorgé, H. Manaa, E. Descroix, P. Laporte, and Y. Guyot, “Excited-state absorption of Tm3+-doped single crystals at photon-avalanche wavelengths,” J. Appl. Phys. 79(8), 4323–4329 (1996). [CrossRef]  

6. S. Kück, L. Fornasiero, E. Mix, and G. Huber, “Excited state absorption and stimulated emission of Nd3+ in crystals. Part I: Y3Al5O12, YAlO3, and Y2O3,” Appl. Phys. B 67(2), 151–156 (1998). [CrossRef]  

7. C. Labbe, J. L. Doualan, S. Girard, R. Moncorgé, and M. Thuau, “Absolute excited state absorption cross section measurements in Er3+:LiYF4 for laser applications around 2.8 µm and 551 nm,” J. Phys.: Condens. Matter 12(30), 6943–6957 (2000). [CrossRef]  

8. S. D. Jackson, “Cross relaxation and energy transfer upconversion processes relevant to the functioning of 2 µm Tm3+-doped silica fibre lasers,” Opt. Commun. 230(1-3), 197–203 (2004). [CrossRef]  

9. A. I. Burshtein, “Hopping mechanism of energy transfer,” Sov. JETP Phys. 35, 882–885 (1972). [CrossRef]  

10. L. Agazzi, K. Worhoff, and M. Pollnau, “Energy-transfer-upconversion models, their applicability and breakdown in the presence of spectroscopically distinct ion classes: A case study in amorphous Al2O3:Er3+,” J. Phys. Chem. C 117(13), 6759–6776 (2013). [CrossRef]  

11. S. A. Pollack, D. B. Chang, and N. L. Moise, “Upconversion-pumped infrared erbium laser,” J. Appl. Phys. 60(12), 4077–4086 (1986). [CrossRef]  

12. T. Komukai, T. Yamamoto, T. Sugawa, and Y. Miyajima, “Upconversion pumped thulium-doped fluoride fiber amplifier and laser operating at 1.47 µm,” IEEE J. Quantum Electron. 31(11), 1880–1889 (1995). [CrossRef]  

13. F. Heine, E. Heumann, T. Danger, T. Schweizer, G. Huber, and B. Chai, “Green upconversion continuous wave Er3+:LiYF4 laser at room temperature,” Appl. Phys. Lett. 65(4), 383–384 (1994). [CrossRef]  

14. T. Hebert, R. Wannemacher, R.M. Macfarlane, and W. Lenth, “Blue continuously pumped upconversion lasing in Tm:YLiF4,” Appl. Phys. Lett. 60(21), 2592–2594 (1992). [CrossRef]  

15. H. Scheife, G. Huber, E. Heumann, S. Bär, and E. Osiac, “Advances in up-conversion lasers based on Er3+ and Pr3+,” Opt. Mater. (Amsterdam, Neth.) 26(4), 365–374 (2004). [CrossRef]  

16. M. F. Joubert, S. Guy, B. Jacquier, and C. Linares, “The photon-avalanche effect: review, model and application,” Opt. Mater. (Amsterdam, Neth.) 4(1), 43–49 (1994). [CrossRef]  

17. J. S. Chivian, W. E. Case, and D. D. Eden, “The photon avalanche: A new phenomenon in Pr3+-based infrared quantum counters,” Appl. Phys. Lett. 35(2), 124–125 (1979). [CrossRef]  

18. M.E. Koch, A.W. Kueny, and W. E. Case, “Photon avalanche upconversion laser at 644 nm,” Appl. Phys. Lett. 56(12), 1083–1085 (1990). [CrossRef]  

19. W. Lenth and R.M. Macfarlane, “Excitation mechanisms for upconversion lasers,” J. Lumin. 45(1-6), 346–350 (1990). [CrossRef]  

20. L. Guillemot, P. Loiko, R. Soulard, A. Braud, J.L. Doualan, A. Hideur, R. Moncorgé, and P. Camy, “Thulium laser at∼ 2.3 µm based on upconversion pumping,” Opt. Lett. 44(16), 4071–4074 (2019). [CrossRef]  

21. R. C. Stoneman and L. Esterowitz, “Efficient, broadly tunable, laser-pumped Tm:YAG and Tm:YSGG CW lasers,” Opt. Lett. 15(9), 486–488 (1990). [CrossRef]  

22. J.F. Pinto, L. Esterowitz, and G.H. Rosenblatt, “Tm3+:YLF laser continuously tunable between 2.20 and 2.46µm,” Opt. Lett. 19(12), 883–885 (1994). [CrossRef]  

23. P. Loiko and M. Pollnau, “Stochastic model of energy-transfer processes among rare-earth ions. Example of Al2O3: Tm3+,” J. Phys. Chem. C 120(46), 26480–26489 (2016). [CrossRef]  

24. K. van Dalfsen, S. Aravazhi, C. Grivas, S. M. García-Blanco, and M. Pollnau, “Thulium channel waveguide laser with 1.6 W of output power and ∼80% slope efficiency,” Opt. Lett. 39(15), 4380–4383 (2014). [CrossRef]  

25. A. Tyazhev, F. Starecki, S. Cozic, P. Loiko, L. Guillemot, A. Braud, F. Joulain, M. Tang, T. Godin, A. Hideur, and P. Camy, “Watt-level efficient 2.3 µm thulium fluoride fiber laser,” Opt. Lett. 45(20), 5788–5791 (2020). [CrossRef]  

26. J. Y. Allain, M. Monerie, and H. Poignant, “Blue upconversion fluorozirconate fibre laser,” Electron. Lett. 26(3), 166–168 (1990). [CrossRef]  

27. S. G. Grubb, K. W. Bennett, R. S. Cannon, and W. F. Humer, “CW room-temperature blue upconversion fibre laser,” Electron. Lett. 28(13), 1243–1244 (1992). [CrossRef]  

28. G. Qin, S. Huang, Y. Feng, A. Shirakawa, M. Musha, and K. I. Ueda, “Power scaling of Tm3+ doped ZBLAN blue upconversion fiber lasers: modeling and experiments,” Appl. Phys. B 82(1), 65–70 (2006). [CrossRef]  

29. J. W. Szela and J. I. Mackenzie, “Excited-state absorption measurements of Tm3+-doped crystals,” Proc. SPIE 8433, 84331O (2012). [CrossRef]  

30. J. Koetke, T. Jensen, and G. Huber, “Infrared stimulated emission and excited state absorption cross sections of Er3+:YAG and Tm3+:YAG lasers,” in Advanced Solid State Lasers, B. Chai and S. Payne, eds., Vol. 24 of OSA Proceedings Series (Optica Publishing Group, 1995), paper IL5.

31. A. Braud, S. Girard, J. L. Doualan, M. Thuau, R. Moncorgé, and A.M. Tkachuk, “Energy-transfer processes in Yb:Tm-doped KY3F10, LiYF4, and BaY2F8 single crystals for laser operation at 1.5 and 2.3 µm,” Phys. Rev. B 61(8), 5280–5292 (2000). [CrossRef]  

32. P. Le Boulanger, J. L. Doualan, S. Girard, J. Margerie, and R. Moncorgé, “Excited-state absorption spectroscopy of Er3+-doped Y3Al5O12, YVO4, and phosphate glass,” Phys. Rev. B 60(16), 11380–11390 (1999). [CrossRef]  

33. P. Loiko, J. L. Doualan, L. Guillemot, R. Moncorgé, F. Starecki, A. Benayad, E. Dunina, A. Kornienko, L. Fomicheva, A. Braud, and P. Camy, “Emission properties of Tm3+-doped CaF2, KY3F10, LiYF4, LiLuF4 and BaY2F8 crystals at 1.5 µm and 2.3 µm,” J. Lumin. 225, 117279 (2020). [CrossRef]  

34. L. Guillemot, P. Loiko, A. Braud, J.L. Doualan, A. Hideur, M. Koselja, R. Moncorge, and P. Camy, “Continuous-wave Tm:YAlO3 laser at ∼2.3 µm,” Opt. Lett. 44(20), 5077–5080 (2019). [CrossRef]  

35. S. A. Payne, L. L. Chase, L. K. Smith, W. L. Kway, and W. F. Krupke, “Infrared cross-section measurements for crystals doped with Er3+, Tm3+, and Ho3+,” IEEE J. Quantum Electron. 28(11), 2619–2630 (1992). [CrossRef]  

36. M. Dulick, G. E. Faulkner, N. J. Cockroft, and D. C. Nguyen, “Spectroscopy and dynamics of upconversion in Tm3+:YLiF4,” J. Lumin. 48-49, 517–521 (1991). [CrossRef]  

37. A. Lupei, V. Lupei, S. Grecu, C. Tiseanu, and G. Boulon, “Crystal-field levels of Tm3+ in gadolinium gallium garnet,” J. Appl. Phys. 75(9), 4652–4657 (1994). [CrossRef]  

38. A. Braud, “Caractéristiques spectroscopiques et émission laser de l'ion Tm3+ à 1.5 µm dans les fluorures,” PhD dissertation (Université de Caen, 1999).

39. S. Renard, P. Camy, A. Braud, J.L. Doualan, and R. Moncorgé, “CaF2 doped with Tm3+: A cluster model,” J. Alloys Compd. 451(1-2), 71–73 (2008). [CrossRef]  

40. L. Guillemot, P. Loiko, R. Soulard, A. Braud, J.L. Doualan, A. Hideur, and P. Camy, “Close look on cubic Tm: KY3F10 crystal for highly efficient lasing on the 3H43H5 transition,” Opt. Express 28(3), 3451–3463 (2020). [CrossRef]  

41. S. So, J.L. Mackenzie, D.P. Shepherd, W.A. Clarkson, J.G. Betterton, and E.K. Gorton, “A power-scaling strategy for longitudinally diode-pumped Tm:YLF lasers,” App. Phys. B 84(3), 389–393 (2006). [CrossRef]  

42. P. Loiko, R. Soulard, L. Guillemot, G. Brasse, J.L. Doualan, A. Braud, A. Tyazhev, A. Hideur, F. Druon, and P. Camy, “Efficient Tm:LiYF4 lasers at 2.3 µm: Effect of energy-transfer upconversion,” IEEE J. Quantum Electron. 55(6), 1700212 (2019). [CrossRef]  

43. S. A. Payne, L. K. Smith, W. L. Kway, J. B. Tassano, and W. F. Krupke, “The mechanism of Tm to Ho energy transfer in LiYF4,” J. Phys.: Condens. Matter 4(44), 8525–8542 (1992). [CrossRef]  

44. S. A. Payne, G. D. Wilke, L. K. Smith, and W. F. Krupke, “Auger upconversion losses in Nd-doped laser glasses,” Opt. Commun. 111(3-4), 263–268 (1994). [CrossRef]  

45. F. Auzel, “Multiphonon-assisted anti-Stokes and Stokes fluorescence of triply ionized rare-earth ions,” Phys. Rev. B 13(7), 2809–2817 (1976). [CrossRef]  

46. P. Loiko, E. Kifle, L. Guillemot, J.L. Doualan, F. Starecki, A. Braud, M. Aguiló, F. Díaz, V. Petrov, X. Mateos, and P. Camy, “Highly efficient 2.3 µm thulium lasers based on a high-phonon-energy crystal: evidence of vibronic-assisted emissions,” J. Opt. Soc. Am. B 38(2), 482–495 (2021). [CrossRef]  

47. M. Falconieri, A. Lanzi, G. Salvetti, and A. Toncelli, “Fluorescence dynamics in Tm,Ho:YLF following 800 nm pulsed laser excitation,” Appl. Phys. B 66(2), 153–162 (1998). [CrossRef]  

48. B. M. Walsh, N. P. Barnes, M. Petros, J. Yu, and U. N. Singh, “Spectroscopy and modeling of solid state lanthanide lasers: Application to trivalent Tm3+ and Ho3+ in YLiF4 and LuLiF4,” J. Appl. Phys. 95(7), 3255–3271 (2004). [CrossRef]  

49. E. Y. Perlin, A. M. Tkachuk, M. J. Joubert, and R. Moncorge, “Cascade-avalanche up-conversion in Tm3+:YLF crystals,” Opt. Spectrosc. 90(5), 691–700 (2001). [CrossRef]  

50. P. Loiko, R. Soulard, G. Brasse, J.L. Doualan, B. Guichardaz, A. Braud, A. Tyazhev, A. Hideur, and P. Camy, “Watt-level Tm:LiYF4 channel waveguide laser produced by diamond saw dicing,” Opt. Express 26(19), 24653–24662 (2018). [CrossRef]  

51. R. Soulard, M. Salhi, G. Brasse, P. Loiko, J.L. Doualan, L. Guillemot, A. Braud, A. Tyazhev, A. Hideur, and P. Camy, “Laser operation of highly-doped Tm:LiYF4 epitaxies: towards thin-disk lasers,” Opt. Express 27(6), 9287–9301 (2019). [CrossRef]  

Data Availability

Data underlying the results presented in this paper are not publicly available at this time but may be obtained from the authors upon reasonable request.

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 (11)

Fig. 1.
Fig. 1. Two-photon processes: (a) excited-state absorption, (b) energy-transfer upconversion and (c) two-photon absorption. |0>, ground-state, |1 > and |2>, excited-states, |1'>, virtual state.
Fig. 2.
Fig. 2. Photon avalanche effect scheme: GSA and ESA – ground- and excited-state absorption, respectively, CR – cross-relaxation, -ph and + ph – absorption / generation of phonons.
Fig. 3.
Fig. 3. Thulium ions: (a) simplified energy-level scheme showing laser transitions in the near- and mid-infrared; (b-c) pumping schemes of Tm3+ ions: (b) conventional pumping; (c) upconversion pumping, GSA and ESA – ground- and excited-state absorption, CR - cross-relaxation, NR – multiphonon non-radiative relaxation.
Fig. 4.
Fig. 4. Summary of excited-state absorption transitions of thulium ions exploited so far (bold arrows – this work), the corresponding ground-state absorption is also shown.
Fig. 5.
Fig. 5. (a) Scheme of the pump-probe setup used for ESA measurements: AOM, acousto-optic modulator, M1 and M2, folding mirrors, L1, focusing lens, L2 – L4, wide-aperture lenses, νpump and νprobe, pump and probe modulation frequencies, respectively, P, polarizer, F, filter; (b) Energy-level scheme of Tm3+ ions illustrating the principle of the pump-probe method.
Fig. 6.
Fig. 6. Evaluation of the ESA cross-section spectra, σESA, for the Tm:LiYF4 crystal in the near-infrared: violet – calibrated raw spectrum ∼ΔI/I, blackσGSA spectra, greenσSE spectra, red - σESA spectra. The light polarization is π.
Fig. 7.
Fig. 7. Interpretation of the polarized ESA cross-section, σESA, spectra for Tm3+ ions in the LiYF4 crystal in the near-infrared: (a,c) the 3F43F2,3 transitions, (b,d) the 3F43H4 transition, light polarizations are (a,b) π and (c,d) σ. Curves – measured spectra, in (c,d), spectra calculated using the reciprocity method (RM) are shown for comparison, vertical dashes – ED electronic transitions of Tm3+ ions according to the polarization selection rules.
Fig. 8.
Fig. 8. Polarized (where applicable) ESA cross-section, σESA, spectra for Tm3+ ions in fluoride materials: (a, b) isotropic hosts: ZBLAN glass, cubic KY3F10 and CaF2 crystals; (c-f) uniaxial crystals: (c,d) tetragonal LiYF4 and (e,f) LiLuF4 and (g,h) biaxial crystal: monoclinic BaY2F8. Transitions: (a,c,e,g) 3F43F2,3 and (b,d,f,h) 3F43H4. The spectral resolution is indicated on the graphs (it is the same in left and right panels).
Fig. 9.
Fig. 9. Polarized (where applicable) ESA cross-section, σESA, spectra for Tm3+ ions in oxide crystals: (a,b) cubic Y3Al5O12 and (c,d) orthorhombic YAlO3 crystals. Transitions: (a,c) 3F43F2,3 and (b,d) 3F43H4. The spectral resolution is indicated on the graphs.
Fig. 10.
Fig. 10. Energy-transfer upconversion (ETU), 3F4 + 3F43H4 + 3H6, for Tm3+ ions. Two elementary processes are shown: (a) energy-migration and (b) phonon-assisted “direct” ETU.
Fig. 11.
Fig. 11. Evaluation of the ETU parameter for Tm:LiYF4: (a) polarization-averaged SE (3F43H6), GSA (3H63F4) and ESA (3F43H4) spectra plotted in semi-log scale, solid curves – measured spectra with phonon sidebands calculated using Eq. (7); (b) summary of the ETU parameters KETU reported so far: symbols – literature data, line – this work, calculation using Eqs. (4)–(6).

Tables (4)

Tables Icon

Table 1. Summary of the Studied Thulium-Doped Materials

Tables Icon

Table 2. Crystal-Field Splitting of Selected Tm3+ Multiplets in LiYF4

Tables Icon

Table 3. Polarization Selection Rules for Electric- and Magnetic-Dipole Transitions in S4 Sites

Tables Icon

Table 4. Excited-State Absorption Peaks* of Tm3+ Ions in Fluoride and Oxide Materials Suitable for Upconversion Pumping at ∼1 µm

Equations (9)

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

Δ I I ( λ ) = A N L ( σ G S A ( λ ) + σ S E ( λ ) σ E S A ( λ ) ) ,
σ E S A ( λ ) = σ S E ( λ ) Z 2 Z 1 exp [ ( h c / λ ) E Z P L k T ] ,
Z m = k g k m e E k m k T ,
C D D = 3 c 8 π 4 n 2 σ S E ( λ ) σ G S A ( λ ) d λ ,
C D A = 3 c 8 π 4 n 2 σ E S A ( λ ) σ S E ( λ ) d λ ,
W E T U = π ( 2 π 3 ) 5 / 2 C D D C D A N T m 2 ,
W E T U = K E T U N T m = C E T U N T m 2 ,
σ S = σ 0 exp ( α S Δ E ) ,
σ AS = σ 0 exp ( α AS Δ E ) ,
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.