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

Insight into neodymium clusters and spectroscopic properties in CaF2 single crystals

Open Access Open Access

Abstract

In this work, the relationship between local structures and the photoluminescence parameters of neodymium activated fluorite crystals has been preliminary investigated in CaF2 crystals. The first principles calculation results reveal that as the size of codoped rare earth ions become smaller, the corresponding stable clusters twists progressively from cubic sublattice to square antiprism. The regulated spectral line properties and the concentration dependent photoluminescence parameters, including the absorption, emission and decays, agree well with the calculated results. In detail, the cubic stable clusters of Nd3+,La3+:CaF2 and Nd3+,Gd3+:CaF2 crystals possess similar spectral characteristics, while the other category of Nd3+,Y3+:CaF2 and Nd3+,Lu3+:CaF2 resembles both in the square antiprism cluster structures and spectra. Besides, a crystal with ultra long lifetime near 600 µs, wide emission bandwidth and high thermal conductivity has been obtained, which is promising for generation of the LD-pumped ultrafast lasers. By connecting the enriched spectral properties with local structures, it is highly expected to design new laser materials through regulating the local structures of active ions in rare earth doped fluorite crystals.

© 2023 Optica Publishing Group under the terms of the Optica Open Access Publishing Agreement

1. Introduction

Ultrafast and high power lasers have attracted great interests due to their wide applications in fusion energy, high energy nuclear physics and attosecond chemistry [1], et al. The ultrafast and ultraintense lasers have been moving towards all-solid-state, high repetition rate, and new wavelength bands [2]. Presently, the ultrafast and ultraintense laser technologies in the near infrared bands are the most mature ones. However, the poor thermal conductivities in the conventionally adopted laser gain materials made the laser repetition rate be low even in the near infrared bands, hence new laser materials were essentially needed to develop high repetition frequency pulses.

Recently, the neodymium doped calcium fluoride crystals have played a more crucial role in ultrafast and high power lasers. The crystals have excellent performance for both the high thermal conductivity (10 W·m-1·K-1) and the broad photoluminescence band (20∼30 nm). Besides, they exhibit low phonon energy, moderate emission cross-section (3∼6 × 10−20 cm2), high upper level lifetime (300-480 µs) and small nonlinear refractive index as low as 0.43 esu [35]. Qin et al. achieved the shortest mode-locked pulses of 103 fs by Nd,Y:CaF2 crystal, which was the first time to realize 100-fs level pulses in Nd-doped laser crystals [6]. The maximum output pulse energy of 5.1 mJ was accomplished with 5 Hz repetition in Nd,Y:SrF2 crystals [7]. The Nd,Y:CaF2 crystals was available for the gain medium of Inertial Confinement Fusion (ICF) laser drivers and a small signal gain of 2.7 is achieved which was nearly twice as that of Nd-glass [8]. These results indicated clearly that Nd-doped alkaline earth fluoride laser crystals are promising candidates for generation of high repetition frequency, ultrafast and high power lasers.

In fact, Nd:CaF2 crystal was studied shortly after the first realization of ruby lasers, and the laser resonant was greatly limited due to the serious luminescence quenching. In order to break the coherent dipolar-dipolar interactions among Nd3+ ions, Kaminskii et al. firstly introduced La3+, Gd3+, Y3+, Lu3+ in Nd3+-doped fluoride crystals, then the photoluminescence intensity was improved significantly [912]. With deeper research, the line characteristics of the spectra were found to be varied with the codopant species, which was supposed to be connected with the rare earth clusters [3,1315]. Catlow et al. reported that the rare earth ions tend to agglomerate together, then dimers, trimers and even hexamers could be formed in the crystals by means of empirical potential simulation [16]. Subsequently, Su et al. investigated the rare earth doped fluorides by using the density functional theory based first principles calculation [17,18]. The cluster features of the rare earth ions were found to be different, which depended on the rare earth ions and matrix crystals, as well as the doping concentrations [1921]. They demonstrated the rare earth clusters in SrF2 crystal as an example, which showed that the cluster structures relied on the size of rare earth ions and evolved with the doping concentrations [22].

Compared with SrF2, the initial clustering concentration in CaF2 crystals was two orders of magnitude lower [23,24]. The spectral properties in CaF2 crystals were also different. Since the rare earth clusters are intimately related with the spectroscopic properties [2527], however, there are rare reports on how the codopant species impact the cluster structure and the spectra properties of the crystals. By systematic research, this work focus on the clusters and spectra of La3+/Gd3+/Y3+/Lu3+ codoped Nd:CaF2. The results show that the cluster structures varied with the codoped ions, which contribute to the differed spectral properties.

2. Methodology

2.1 First principles calculation

The calculations of trivalent rare earth clusters in CaF2 crystal were carried out by using the VASP codes [28,29]. The projector augmented wave (PAW) potential was utilized. The formalism for the calculations is related to the Perdew-Burke-Ernzerhof (PBE) function within the generalized gradient approximation (GGA) [30,31]. And the spin polarization was included in all calculations. The Ca 3p64s2, F 2s22p5, La 5s25p65d16s2, Nd 5s25p65d16s2, Gd 5s25p65d16s2, Y 4s24p64d25s1, Lu 5s25p65d16s2 states were considered as valence electrons. Due to interactions among the ions, the plane wave basis has set the cut off energy as 550 eV. To obtain sufficient accuracy, the 2 × 2 × 2 supercell and 1 × 1 × 1 Gamma k-meshes were considered. The converged energy gap was 10−5 eV and force on individual atoms reached 0.01 eV·Å−1.

The symbol i1i2|v|d|sr-O/A/T/Cn was applied to describe the rare earth clusters based on Refs. [16,20]. i1 and i2 represent the number of Nd and second kind of codoped rare earth atoms, respectively; v is the number of lattice fluorine vacancies; d is the number of lattice fluorines that deviated from the normal site; and s means the number of interstitial fluorines (Fi-); r = 1 and 2 point to that the interstitial fluorines located at the nearest and next nearest site of rare earth ion, respectively; O, A, T, C respectively refer to the shape of the cluster resembling a circle (O), an armchair (A in abbreviation), a crescent moon (C) and a tack nail (T), respectively. The suffix n means the serial number to supplement the insufficient notation. The formation energy of the cluster was calculated by Eq. (1):

$$\begin{aligned}&\Delta E = \left( {E_{{\rm tot}} + E_0} \right)-i_1\cdot E_1-i_2\cdot E_2-\left( {s-v} \right)\cdot E_3-\{ \left[ {i_1 + i_2 + \left( {s-v} \right)} \right]-\\ &\quad\quad\quad\quad\quad\quad{[{{i_1} + {i_2} + ({s - v} )} ]^2}\} \cdot{E_{\textrm{corr}}}\end{aligned}$$
Where Etot means the total relaxed energy of the clusters; E0 is the energy of pure CaF2 crystal; E1, E2 and E3 respectively point to the relaxed energy of 1102|0|0|0, 0112|0|0|0 and 0102|0|0|1 center in a 2 × 2 × 2 supercell; Ecorr denotes the potential alignment and image charge corrections obtained by Eq. (2) [32,33]:
$${E_{\textrm{corr}}} = ({1 + g} )\frac{{{q^2}\alpha }}{{2\varepsilon L}}$$

In Eq. (2), g is the scaling factor of -0.34 for face-centered cubic structure [34]; q is the net charge, α is the Madelung constant of 5.0388 for fluorite [35]; ε and L respectively denote to the static dielectric constant and supercell dimension (2 × 2 × 2), ε equals to 6.812 at 300 K [36], and L equals to 10.9 Å. Hence for a 2 × 2 × 2 supercell, the correction energy of CaF2 with a net charge was 0.323 eV [37].

The partial Density of States (pDOS) calculations were performed based on the optimized structures. The 4 × 4 × 4 Gamma k-meshes was adopted to sampling the first Brillouin zone. The LSDA + U method by Dudarev was considered for pseudopotential corrections and Coulomb onsite interactions [38]. In the work U = 0.5, 6 and 1 were respectively utilized for lanthanum, neodymium and yttrium. It is noted that the Nd 4f35s25p65d16s2 with 4f electrons unfreezing was adopted in the the Partial Density of States(pDOS).

2.2 Sample preparation and characterization

Four groups of single crystals Nd3+,R3+:CaF2 (R = La, Gd, Y, Lu) were grown by temperature gradient technique method. The concentrations of Nd3+ is around 0.5 at.% measured by Inductive Coupled Plasma Optical Emission Spectrometer (ICP-OES). The doping concentrations of R3+ were 0.5, 2, 5 and 8 at.% in four classes of crystals. The sample of 0.5% Nd3+,8%Ce3+:CaF2 was also grown as a comparison with other crystals. 1 wt.% PbF2 was added as oxygen scavenger. Raw materials of NdF3, LaF3, GdF3, YF3, LuF3, CaF2 and PbF2 were used with purity of 99.99%. All raw materials were weighted stoichiometrically and well-mixed, and then loaded into a graphite crucible. The growth program wasn’t started to run until the vacuum reached 10−3 Pa. The crystal growth was followed with rate of 1.5 K/h, as described in Ref. [39]. After the growth process, the as grown crystals were prepared with dimension of Φ11 × 2 mm3 for spectral measurement. The samples were polished double faces.

Room temperature absorption spectra were tested by Cary 5000 UV-Vis-NIR Spectrophotometer, with step of 1 nm, spectral band width (SBW) of 2 nm and dwell time of 0.1 s. The fluorescence spectra were recorded by FLS980 time-resolved fluorimeter with a thermoelectric (TE) cooled InGaAs detector. The fluorescence decay curves were measured by using a TDS 3052 oscilloscope following excitation at 794 nm by Xenon flashlamp. In addition, the detection wavelength for the decay lifetime measurement is 1060 nm. If the curves are nonexponential, then the following equation is used to calculate the average lifetime.

$${\tau _{em}} = \frac{{\mathop \smallint \nolimits_{t = 0}^\infty I(t )tdt}}{{\mathop \smallint \nolimits_{t = 0}^\infty I(t )dt}}$$

Based on these parameters, the peak emission cross sections are calculated by Eq. (4).

$${\sigma _{em}} = \frac{{\lambda _{peak}^4}}{{8\pi c{n^2}}}\cdot \frac{{{\beta _2}\cdot \eta }}{{{\tau _{em}}\Delta {\lambda _{em}}}}$$
where λpeak means the peak emission line of 4F3/24I11/2, β2 and Δλem point to the branching ratio and emission bandwidth of 4F3/24I11/2, respectively.

3. Results and discussion

3.1 Rare earth clusters simulation

When the rare earth ions are introduced in CaF2, it will replace the Ca2+ and interstitial Fi- work as charge compensation. There are several nonequivalent sites for the charge compensation Fi- to occupy. Then the 11|0|0|11 and 11|0|0|12 with Fi- at the nearest and next nearest sites of rare earth ions respectively are emerged. These monomers would be further aggregated to be clusters and about twenty kinds of rare earth centers are observed in Er3+:CaF2 [40]. Herein, the rare-earth clusters in CaF2 crystal are modeled based on Ref. [21], and then totally relaxed. The potential thermodynamic stable clusters are shown in Fig. 1 and Table 1. The thermodynamic stable centers are classified into three stages. The first stage consists of monomers containing only one rare earth ions; the second stage consists of clusters with more than one rare earth ions and the ionic sublattice within the cluster is remained to be cubic, and the third stage points to square antiprism ionic sublattice. It can be seen in Fig. 1 that, from La3+, Nd3+, Gd3+, Y3+ to Lu3+, formation energy of the first stage monomer center 11|0|0|11 increases while that of 11|0|0|12 decreases. It means that the charge compensation Fi- tends to occupy the next nearest site of rare earth ions as the ionic radius decreases. At the same time, one more lattice fluorine departs from the normal site in the second stage dimers 21|0|1|21 and 21|0|1|31 of large size La3+ ion clusters. Then 21|0|2|21 and 21|0|2|31 centers were generated for small size ion such as Y3+. The cubic center 41|1|0|41 is thermodynamic stable in large size La3+ and Nd3+ doped CaF2, and it becomes unstable for Y3+ and Lu3+. There are two lattice fluorines deviating in the cubic cluster 31|0|2|41-T for Nd3+ and Gd3+. While for Y3+ and Lu3+, five and six more normal fluorines relaxed, respectively. More importantly, the ionic sublattice varied from cubic to square antiprism structure. As for the third stage clusters, there shows no stable centers in large radius ions like La3+ and Nd3+ and the stability was significantly improved in small size ions Lu3+ and Y3+. As to Gd3+, the number of cubic stable clusters account for a large propotion while there is only a small percentage stable centers in the third stage. These results clearly suggest that the characters of centers change with the radius of rare-earth ions. The second stage cubic sublattice clusters are more stable in large size ions La3+, Nd3+ and Gd3+, while the third stage square antiprism clusters are more stable for small size ions Y3+ and Lu3+. Accordingly, cluster characters of La3+, Nd3+ and Gd3+ were classified as one category and that of Y3+ and Lu3+ as another category.

 figure: Fig. 1.

Fig. 1. The thermodynamic stable clusters in La3+-, Nd3+-, Gd3+-, Y3+-, Lu3+-doped CaF2 crystals, these centers are divided into three groups, “1”, “2” and “3” represent the cubic monomers, cubic sublattice clusters and square antiprism structure clusters, respectively.

Download Full Size | PDF

Tables Icon

Table 1. Formation energy of the thermodynamic stable rare earth clusters in CaF2 crystal.

The number of rare earth ions dependent formation energy is also depicted. It can be seen in Fig. 2 that, as the number of rare earth ions within the same doping center increases, the formation energy reduces linearly. It indicates that the clusters would be more stable when the rare earth ions agglomerate together, which is in line with the reported results [37]. The energy slopes for La3+, Nd3+, Gd3+, Y3+ and Lu3+ are -1.54, -1.54, -1.71, -1.61 and -1.74 eV, respectively. As the formation slopes become steeper, the stabilities of clusters are enhanced. Thus the magnitude of energy slopes reflects the difficulty of aggregating. Hence the smaller values observed in small size rare earth ions mean, that it is easier for not only the process of aggregating, but also the transforming of ionic sublattice from cubic to square antiprism structure. Subsequently, the lattice mismatch was significantly alleviated. Thus the formation energy and slopes are further reduced for smaller size rare-earth ions Y3+ and Lu3+.

 figure: Fig. 2.

Fig. 2. The number of rare earth ions dependent formation energy of the thermodynamic stable centers.

Download Full Size | PDF

The rare earth cluster structures in singly-doped calcium fluorite are already figured out. However, clusters in the rare earth codoped crystals remain unclear. Since the codoped fluorites, especially Nd3+ and R3+ codoped crystals are potential candidates for generation of high repetition rate high power lasers, the cluster structures in the codoped crystals are also vital. Referring to the characters of rare earth clusters stated above, it was assumed that the structures of mixed [Nd3+-La3+] are similar with that of [Nd3+-Gd3+], and of [Nd3+-Y3+] resemble that of [37] [37]. The centers of [Nd3+-La3+] and [Nd3+-Y3+] were therefore chosen in order to simplify the calculation. The mixed clusters were then modeled and relaxed. The thermodynamic stable clusters are shown in Fig. 3 and Tables 23. As can be seen that only the cubic sublattice centers are thermodynamically stable in [Nd3+-La3+] mixed clusters. The structure characters are similar with that of Nd3+ or La3+ singly-doped crystals in which the cubic sublattice centers are stable. Compared with [Nd3+-La3+], however, the proportion of cubic sublattice centers decrease in the Nd3+ and Y3+ codoped crystals. There is quite amount of cubic sublattice [Nd3+-Y3+] clusters were converted to square antiprism structure centers. It is in line with the structure characters of Y3+-doped CaF2. Therefore, the mixed cluster structures depend on the compositions and structural features of the rare earth centers.

 figure: Fig. 3.

Fig. 3. Thermodynamic stable centers of the mixed [Nd3+-La3+] and [Nd3+-Y3+] clusters in CaF2 crystal.

Download Full Size | PDF

Tables Icon

Table 2. The formation energy (FE), averaged bond length (ABL), coordination number (CN) and the ionic sublattice (SBL) around Nd3+ in [Nd3+-La3+] centers, the SAP means square antiprism sublattice.

Tables Icon

Table 3. The formation energy (FE), averaged bond length (ABL), coordination number (CN) and the ionic sublattice (SBL) around Nd3+ in [Nd3+-Y3+] centers, the SAP means square antiprism sublattice.

According to Tables 23, the number of rare earth ions dependent formation energy is illustrated. As for [Nd3+-La3+] clusters (m = 0 ∼ 4), the formation energy decreases linearly with the number of La3+ atoms rising, as the solid lines shown from Fig. 4. The slopes of clusters with m = 0, 1 and 2 are approximately the same, and the value is around -1.56 eV, agrees well with that of La3+ or Nd3+ singly doped crystals. Besides, as the dotted lines show, the formation energy of the [Nd3+-La3+] clusters are dropped when compared with that of pure Nd3+ or La3+ centers. These are strong evidence that the introduction of La3+ into Nd3+ clusters contribute to a more stable framework and two possible reasons should be responsible. First, the different radius sizes of La3+ and Nd3+ ions would probably bring a large reduction of lattice mismatch. Second, the centers of Nd3+ or La3+ singly doped crystals are thermodynamic stable with cubic sublattice structures which are also expected in Nd3+ and La3+ codoped sample.

 figure: Fig. 4.

Fig. 4. Number of R3+ ions dependent formation energy of the mixed [Nd3+-R3+] clusters in CaF2 crystal, R = La (a) and Y (b), respectively, the m and n represent number of Nd3+ and R3+ atoms in the clusters, respectively.

Download Full Size | PDF

The formation energy of [Nd3+-Y3+] centers is presented as a comparison. It can be seen from the solid lines, that formation energies are also reduced linearly. The slope values are similar to that of pure Y3+ centers, especially in the centers with more Y3+ ions. The lower slope of [Nd3+-Y3+] centers than that of [Nd3+-La3+] indicate that it is easier to be clustering for Nd3+ and Y3+ [37]. For the dotted lines, formation energy of the cubic sublattice centers with m + n = 2 and 3 increase as the number of Y3+ boosts up, while that of square antiprism structure clusters with m + n = 4, 5 and 6 slightly decrease. These results clearly demonstrate that stabilities of the cubic structure centers are weakened and they tend to shift towards square antiprism centers with addition of Y3+. The cluster characters of La3+ and Y3+ are different, which then contributes to the varied structure characteristics of [Nd3+-La3+] and [Nd3+-Y3+] mixed centers.

The bonding characters are further analyzed to confirm the varied cluster structures. Given that the relative stabilities of 11|0|0|11 and 11|0|0|12 corresponding to that of second and third stage centers depend on the 11|0|0|11 [17], the pDOS of 11|0|0|11 center in La3+, Nd3+ and Y3+ doped CaF2 is therefore calculated. As is shown in Fig. 5(a), the f orbital of La3+ locates at 6.5 eV in the conduction band, which is the lowest when compared with that of f states in Nd3+ and d states in Y3+. It would be easier for the nonbonding charges of interstitial Fi- moving towards the lowest unoccupied states in La3+. The connection between La3+ and Fi- was then strengthened. For Nd3+, although the energy of 4f states in conduction band is higher than that of d states in Y3+, however, the much more intense bonding was observed near the top of the valence band. In terms of interstitial Fi-, as can be seen in Fig. 5(b) that the intensity of 2p states at about 0 eV in Y3+ is the highest. It means that there are more charges located at the Fi-, which would give a strong repulsion to the lattice fluorines around. The cubic sublattice then became unstable and it alters to be square antiprism structure. The bonding characters match well with the structure characters of rare earth centers and the mixed.

 figure: Fig. 5.

Fig. 5. pDOS of the lanthanide and interstitial fluorine in 11|0|0|11 center of CaF2. (a) The states of La3+ (4f), Nd3+ (4f) and Y3+ (4d); (b) the 2p states of the Fi-.

Download Full Size | PDF

3.2 Spectral properties of Nd3+,R3+:CaF2 crystals

On the basis of the aforementioned results, the cubic sublattice clusters prefer to be formed in Nd3+,La3+:CaF2 and Nd3+,Gd3+:CaF2 crystals. While as the codopant changes to be Y3+ and Lu3+, the square antiprism clusters are emerged. With the results in mind, the spectral properties of Nd3+,R3+:CaF2 (R = La, Gd, Y, Lu) crystals were compared to confirm the viewpoint.

The room temperature absorption and emission spectra as well as decay curves of Nd3+,R3+:CaF2 are presented in Fig. 6 and the spectra are all normalized for comparison. Since the clustering of the codopant would be more expressed in the mixed centers when the concentration of R3+ is substantially higher than that of Nd3+, the crystals codoped with 8% R3+ were selected. As can be seen in Fig. 6(a) that, there are two absorptions peaking at 790 and 796 nm, which corresponds to the 4I9/24F5/2 + 2H9/2 transition. When using La3+ and Gd3+ as the codopant, the highest peak emerges at 790 nm while the intensity of 796 nm absorption is relatively weak, which agrees well with the crystal codoped with large size Ce3+. However, the absorption intensity of 796 nm is much stronger than that at 790 nm as codopants switch to be Y3+ and Lu3+ [13].

 figure: Fig. 6.

Fig. 6. Comparison of area normalized absorption (a), intensity normalized emission spectra (b) and (c) decay curves of Nd3+,R3+:CaF2 crystal.

Download Full Size | PDF

As displayed in Fig. 6(b), the emission spectra are also illustrated. The emission band corresponds to the 4F3/24I11/2 transition. The line characters of emission spectra in Nd3+,La3+/Ce3+/Gd3+:CaF2 crystals are quite similar and two peaks are observed. The stronger emission is at 1065 nm and the weaker is at 1049 nm. When Y3+ and Lu3+ are introduced in the crystals, the emissions at 1049 and 1065 nm disappear and the intensity at 1054 nm boosts up. Hence, as additional information, the spectrum of Nd3+,Ce3+:CaF2 crystal further demonstrates that similar spectrum are related to the size of ion radius. In addition, the decay curve of Nd3+,La3+:CaF2 is nearly the same with Nd3+,Gd3+:CaF2, and that of Nd3+,Y3+:CaF2 and Nd3+,Lu3+:CaF2 resemble with each other in Fig. 6(c). Specifically, the decay values of Nd3+,La3+/Gd3+:CaF2 crystals are 481.7 and 566.5 µs, which is relatively high. While the lifetime values of Nd3+,Y3+/Lu3+:CaF2 crystals are 401.0 and 362.7 µs, which are almost the same and shorter. The results of emissions and decays agree well with the absorption spectra. Combined with the calculated results, the absorption of 790 nm and the emissions of 1049 and 1064 nm could be connected with the cubic sublattice centers. While the 796 nm absorption and the 1054 nm emission are bounded with the square antiprism clusters [41].

Furthermore, the codopant concentrations dependent spectral parameters of Nd3+,R3+:CaF2 crystals are summarized and depicted in Fig. 7. As for the quantum efficiencies, they are all improved when the doping concentration increases in Fig. 7(a). The quantum efficiencies are related with the broken of [Nd3+-Nd3+] clusters [15]. As discussed before, the introduced R3+ ions would agglomerate with Nd3+ efficiently as the amount of [Nd3+-R3+] nonquenching clusters increases, which contributes to the enhanced quantum efficiencies. Additionally, the quantum efficiency is connected with the number of photons of absorption and emission. While the local structures impact both the characteristics of absorption and emission spectrum, thus the effects might be neutralized. In respect of the lifetimes, there is a positive correlation between the values and the codopant concentrations. In particular, the lifetimes of Nd3+,La3+:CaF2 and Nd3+,Gd3+:CaF2 crystals are higher than Nd3+,Y3+:CaF2 and Nd3+,Lu3+:CaF2. The variation trends of the concentration dependent lifetimes in Nd3+,La3+:CaF2 and Nd3+,Gd3+:CaF2 exhibited to be the same in Fig. 7(b). As can be seen the trends in Nd3+,Y3+:CaF2 and Nd3+,Lu3+:CaF2 are also approximately the same. It is similar with the concentrations dependent emission cross sections of the codoped crystals in Fig. 7(c). Therefore, the varied photoluminescence parameters of Nd3+,R3+:CaF2 crystals are in accord with the altered rare earth cluster structures.

 figure: Fig. 7.

Fig. 7. The codopant concentrations dependent photoluminescence parameters of Nd3+,R3+:CaF2 crystals, (a) the quantum efficiencies, (b) the fluorescence lifetimes, (c) the peak emission cross sections.

Download Full Size | PDF

It should be noted here that the longer lifetimes and smaller emission cross sections were collected in Nd3+,La3+:CaF2 when compared with that of Nd3+,Y3+:CaF2. The averaged bond length, coordination number and polyhedrons around the active Nd3+ in the mixed clusters were collected in Tables 2 and 3. It shows that a wider span of averaged bond lengths were observed in Nd3+,La3+:CaF2 crystals. While for Nd3+,Y3+:CaF2, span of the averaged bond lengths declines. As can be seen the averaged bond lengths range from 2.38 to 2.49 Å in the La3+-codoped crystal. Similar spanning was observed in the cubic sublattice centers of Nd3+,Y3+:CaF2 crystal. However, it altered to be spanning 2.39 to 2.41 Å in the high order clusters. The longer averaged bond lengths, dispersal distribution of bond lengths and larger coordination numbers as well as the higher symmetric cubic polyhedrons lead to the longer lifetimes and smaller emission cross sections. When the codoped rare-earth ions alter from La3+, Gd3+, Y3+ to Lu3+, the cluster structures shift from cubic to square antiprism, attributing to the shorter lifetime and larger emission cross section. The relationship of cluster structures and photoluminescence parameters were then depicted in Fig. 8. As the local structures varied from cubic to square antiprism, the emission cross sections were improved and simultaneously the lifetimes decrease. The relationship provides a methodology for designing new laser materials by regulating the local structures of rare earth ions.

 figure: Fig. 8.

Fig. 8. The local structures dependent peak emission cross sections and lifetimes in 0.5%Nd3+,8%R3+:CaF2 crystals.

Download Full Size | PDF

The photoluminescence parameters and thermal conductivities of Nd3+,R3+:CaF2 crystals were also listed and compared with other famous Nd3+-doped laser materials, such as Nd:YAG, Nd:YVO4, Nd:YLF, Nd:Ca3La2(BO3)4 and Nd:glass. It can be seen in Table 4, that the bandwidths of Nd,R:CaF2 crystals are comparable with Nd:Ca3La2(BO3)4 and Nd:glass, and is much broader than that in Nd:YAG, Nd:YVO4 and Nd:YLF. The thermal conductivities in Nd,R:CaF2 are much higher than that in Nd:Ca3La2(BO3)4 and Nd:glass, which supports the high repetition rate ultrafast lasers. More importantly, a lifetime near 600 µs was achieved in Nd3+,Gd3+:CaF2 crystal. The high energy storage ability means that the LD-pumped lasers were highly expected in the crystals.

Tables Icon

Table 4. Comparison of the photoluminescence parameters and thermal conductivities at room temperature of the neodymium-doped laser materials.

4. Conclusion

In summary, we have in this work prepared for four kinds of Nd3+,R3+:CaF2 crystals. As to the simple-doped calculation results, the stabilities of cubic sublattice clusters declined and that of square antiprism clusters strengthened when the codopant varied from La3+, Gd3+, Y3+ to Lu3+. The mixed rare earth clusters calculation indicates, that the cluster characters in Nd3+,La3+:CaF2 and Nd3+,Gd3+:CaF2 crystals are nearly the same and could be classified into one category, correspondingly the Nd3+,Y3+:CaF2 and Nd3+,Lu3+:CaF2 crystals are as another one. The pDOS analysis supplements the reason for different cluster structures from the view of binding characteristics. The local structure dependent spectral properties and the concentration dependent photoluminescence parameters results show that the similar line characters were collected in Nd3+,La3+:CaF2 and Nd3+,Gd3+:CaF2. Subsequently, the Nd3+,Y3+:CaF2 and Nd3+,Lu3+:CaF2 possess almost the same spectral features, which agrees with the calculated results. Accordingly, the 790 and 796 nm absorptions were ascribed to the cubic and square antiprism sublattice centers, respectively. And the cubic sublattice centers result in the emissions at 1049 and 1064 nm, while the square antiprism clusters cause the 1054 nm emission. The higher local symmetries around the active ions in cubic sublattice centers contribute to the longer lifetimes (∼ 600 µs) and moderate emission cross sections (∼ 2.5 × 10−20 cm2) in the crystals codoped with large size rare earth ions. The ultralong lifetime and appropriate emission cross sections make the crystals be potential candidates for generation of laser diode-pumped ultrafast pulses.

Funding

Strategic Priority Program of the Chinese Academy of Sciences (XDA25020313); National Key Research and Development Program of China (2022YFB3605702); National Natural Science Foundation of China (61925508, U2230103); Science and Technology Commission of Shanghai Municipality (22ZR1472000); Young Scientists in Basic Research (YSBR-024); International Partnership Program of Chinese Academy of Sciences (121631KYSB20200039).

Disclosures

The authors declare no conflicts of interests.

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. H. Schwoerer, B. Liesfeld, H. P. Schlenvoigt, K. U. Amthor, and R. Sauerbrey, “Thomson-backscattered X rays from laser-accelerated electrons,” Phys. Rev. Lett. 96(1), 014802 (2006). [CrossRef]  

2. U. Keller, “Recent developments in compact ultrafast lasers,” Nature 424(6950), 831–838 (2003). [CrossRef]  

3. S. Y. Pang, F. K. Ma, H. Yu, X. B. Qian, D. P. Jiang, Y. J. Wu, F. Zhang, J. Liu, J. Y. Xu, and L. B. Su, “Highly efficient continuous-wave laser operation of LD-pumped Nd, Gd: CaF2 and Nd, Y: CaF2 crystals,” Laser Phys. Lett. 15(5), 055802 (2018). [CrossRef]  

4. Y. Guo, S. B. Lu, L. B. Su, C. J. Zhao, H. Zhang, and S. C. Wen, “Z-scan measurement of the nonlinear refractive index of Nd3+, Y3+-codoped CaF2 and SrF2 crystals,” Appl. Opt. 54(4), 953–958 (2015). [CrossRef]  

5. Q. Zhang, L. B. Su, D. P. Jiang, F. K. Ma, Z. P. Qin, G. Q. Xie, J. G. Zheng, Q. H. Deng, W. G. Zheng, and L. J. Qian, “Highly efficient continuous-wave laser operation of laser diode-pumped Nd, Y: CaF2 crystals,” Chin. Opt. Lett. 13(7), 071402 (2015). [CrossRef]  

6. Z. P. Qin, G. Q. Xie, J. Ma, W. Y. Ge, P. Yuan, L. J. Qian, L. B. Su, D. P. Jiang, F. K. Ma, and Q. Zhang, “Generation of 103 fs mode-locked pulses by a gain linewidth-variable Nd,Y:CaF2 disordered crystal,” Opt. Lett. 39(7), 1737–1739 (2014). [CrossRef]  

7. J. C. Chen, Y. J. Peng, Z. X. Zhang, H. P. Su, Y. X. Leng, D. Jiang, F. K. Ma, X. B. Qian, F. Tang, and L. B. Su, “Demonstration of a diode pumped Nd, Y co-doped SrF2 crystal based, high energy chirped pulse amplification laser system,” Opt. Commun. 382, 201–204 (2017). [CrossRef]  

8. X. X. Tang, J. S. Qiu, Z. W. Fan, L. B. Su, and H. C. Wang, “Diode-pumped medium-aperture-size square Nd, Y: CaF2 rod amplifier for Inertial Confinement Fusion laser drivers,” Opt. Mater. 58, 445–448 (2016). [CrossRef]  

9. A. A. Kaminskii, N. R. Agamalyan, G. A. Deniseneo, S. E. Sarkisov, and P. P. Fedorov, “Spectroscopy and laser emission of disordered GdF3 CaF2: Nd3+ trigonal crystals,” Phys. Status Solidi A 70(2), 397–406 (1982). [CrossRef]  

10. A. A. Kaminskii, V. V. Osiko, A. M. Prochorov, and Y. K. Voronko, “Spectral investigation of the stimulated radiation of Nd3+ in CaF2-YF3,” Phys. Lett. 22(4), 419–421 (1966). [CrossRef]  

11. K. S. Bagdasarov, Y. K. Voronko, A. A. Kaminskii, L. V. Krotova, and V. V. Osiko, “Modification of the Optical Properties of CaF2-TR3+ Crystals by Yttrium Impurities,” Phys. Status Solidi B 12(2), 905–912 (1965). [CrossRef]  

12. A. A. Kaminskii, “High-temperature Effects Observed in Stimulated Emission from CaF2 and LaF2 Crystals Activated with Nd3+,” Jetp Lett-Ussr 6, 615 (1967).

13. C. Meroni, A. Braud, J. L. Doualan, C. Maunier, S. Montant, and P. Camy, “Spectroscopy and laser gain measurements of CaF2: Nd, X3+, Z3+ (X, Z = Gd, La, Ce, Y, Lu, Sc) crystals for broadband lasers applications,” J. Lumin. 252, 119336 (2022). [CrossRef]  

14. X. H. Li, Q. Q. Hao, D. P. Jiang, Q. H. Wu, Z. Zhang, Z. H. Zhang, J. Liu, and L. B. Su, “Smooth and flat photoluminescence spectra of Nd3+ active ions in tri-doped CaF2 single crystals,” Opt. Mater. Express 10(3), 704–714 (2020). [CrossRef]  

15. Q. Q. Hao, S. Y. Pang, J. Liu, and L. B. Su, “Tunable and passively Q-switched laser operation of Nd, Lu: CaF2 disordered crystal,” Appl. Opt. 57(22), 6491–6495 (2018). [CrossRef]  

16. J. Corish, C. R. A. Catlow, P. W. M. Jacobs, and S. H. Ong, “Defect aggregation in anion-excess fluorites. Dopant monomers and dimers,” Phys. Rev. B 25(10), 6425–6438 (1982). [CrossRef]  

17. F. K. Ma, F. Su, R. F. Zhou, Y. Y. Ou, L. J. Xie, C. M. Liu, D. P. Jiang, Z. Zhang, Q. H. Wu, and L. B. Su, “The defect aggregation of RE3+ (RE = Y, La∼ Lu) in MF2 (M = Ca, Sr, Ba) fluorites,” Mater. Res. Bull. 125, 110788 (2020). [CrossRef]  

18. F. K. Ma, P. X. Zhang, L. B. Su, H. Yin, Z. Li, Q. Lv, and Z. Q. Chen, “The host driven local structures modulation towards broadband photoluminescence in neodymium-doped fluorite crystal,” Opt. Mater. 119, 111322 (2021). [CrossRef]  

19. A. K. Cheetham, B. E. F. Fender, and M. J. Cooper, “Defect structure of calcium fluoride containing excess anions I. Bragg scattering,” J. Phys. C: Solid State Phys. 4(18), 3107–3121 (1971). [CrossRef]  

20. P. J. Bendall, C. R. A. Catlow, J. Corish, and P. W. M. Jacobs, “Defect aggregation in anion-excess fluorites II. Clusters containing more than two impurity atoms,” J. Solid State Chem. 51(2), 159–169 (1984). [CrossRef]  

21. C. R. A. Catlow, “Defect clusters in doped fluorite crystals,” J. Phys. C: Solid State Phys. 6(4), L64–L70 (1973). [CrossRef]  

22. F. K. Ma, H. Q. Zhou, Q. Y. Tang, L. B. Su, M. J. Zhang, P. X. Zhang, H. Yin, Z. Li, Q. Lv, and Z. Q. Chen, “Clusters modification for tunable photoluminescence in Nd3+:SrF2 crystal,” J. Alloys Compd. 899, 162913 (2022). [CrossRef]  

23. Y. V. Orlovskii, T. T. Basiev, S. A. Abalakin, I. N. Vorob’ev, O. K. Alimov, A. G. Papashvili, and K. K. Pukhov, “Fluorescence quenching of the Nd3+ ions in different optical centers in fluorite-type crystals,” J. Lumin. 76-77, 371–376 (1998). [CrossRef]  

24. Y. V. Orlovskii, V. V. Fedorov, T. T. Basiev, M. Altwein, B. Leu, J. Heber, and S. Mirov, “Nonradiative relaxation and inhomogeneous splitting of aggregated optical centers in the Nd3+-doped CaF2 and SrF2 crystals (FLN and decay study),” J. Lumin. 83-84, 361–366 (1999). [CrossRef]  

25. M. Wu, E. H. Song, Z. T. Chen, S. Ding, S. Ye, J. Zhou, S. Xu, and Q. Y. Zhang, “Single-band red upconversion luminescence of Yb3+–Er3+ via nonequivalent substitution in perovskite KMgF3 nanocrystals,” J. Mater. Chem. C 4(8), 1675–1684 (2016). [CrossRef]  

26. J. L. Wang, E. H. Song, M. Wu, W. B. Dai, S. Ye, and Q. Y. Zhang, “Selectively enhanced up-and down-conversion emissions of Er3+ via Yb3+-Mn2+ dimer sensitizing in spinel MgGa2O4: Er3+, Yb3+, Mn2+,” Mater. Res. Bull. 74, 340–345 (2016). [CrossRef]  

27. J. L. Wang, E. H. Song, M. Wu, W. B. Dai, X. F. Jiang, B. Zhou, and Q. Y. Zhang, “Room-temperature green to orange color-tunable upconversion luminescence from Yb3+/Mn2+ co-doped CaO,” J. Mater. Chem. C 4(42), 10154–10160 (2016). [CrossRef]  

28. G. Kresse and J. Hafner, “Ab initio molecular dynamics for liquid metals,” Phys. Rev. B 47(1), 558–561 (1993). [CrossRef]  

29. G. Kresse and J. Furthmüller, “Efficiency of ab-initio total energy calculations for metals and semiconductors using a plane-wave basis set,” Comput. Mater. Sci. 6(1), 15–50 (1996). [CrossRef]  

30. P. E. Blöchl, “Projector augmented-wave method,” Phys. Rev. B 50(24), 17953–17979 (1994). [CrossRef]  

31. J. P. Perdew, K. Burke, and M. Ernzerhof, “Generalized gradient approximation made simple,” Phys. Rev. Lett. 77(18), 3865–3868 (1996). [CrossRef]  

32. G. Makov and M. C. Payne, “Periodic boundary conditions in ab initio calculations,” Phys. Rev. B 51(7), 4014–4022 (1995). [CrossRef]  

33. C. Persson, Y. J. Zhao, S. Lany, and A. Zunger, “n-type doping of CuIn Se2 and CuGa Se2,” Phys. Rev. B 72(3), 035211 (2005). [CrossRef]  

34. S. Lany and A. Zunger, “Assessment of correction methods for the band-gap problem and for finite-size effects in supercell defect calculations: Case studies for ZnO and GaAs,” Phys. Rev. B 78(23), 235104 (2008). [CrossRef]  

35. A. Jockisch, U. Schroder, F. W. de Wette, and W. Kress, “Relaxation and dynamics of the (111) surfaces of the fluorides CaF2 and SrF2,” J. Phys.: Condens. Matter 5(31), 5401–5410 (1993). [CrossRef]  

36. C. Andeen, D. Schuele, and J. Fontanella, “Pressure and temperature derivatives of the low-frequency dielectric constants of the alkaline-earth fluorides,” Phys. Rev. B 6(2), 591–595 (1972). [CrossRef]  

37. F. K. Ma, D. P. Jiang, Z. Zhang, X. Q. Tian, Q. H. Wu, J. Y. Wang, X. B. Qian, Y. Liu, and L. B. Su, “Tailoring the local lattice distortion of Nd3+ by codoping of Y3+ through first principles calculation for tuning the spectroscopic properties,” Opt. Mater. Express 9(11), 4256–4272 (2019). [CrossRef]  

38. S. L. Dudarev, G. A. Botton, S. Y. Savrasov, C. J. Humphreys, and A. P. Sutton, “Electron-energy-loss spectra and the structural stability of nickel oxide: An LSDA+ U study,” Phys. Rev. B 57(3), 1505–1509 (1998). [CrossRef]  

39. Y. Z. Zhou, “Growth of high quality large Nd: YAG crystals by temperature gradient technique (TGT),” J. Cryst. Growth 78(1), 31–35 (1986). [CrossRef]  

40. D. R. Tallant and J. C. Wright, “Selective laser excitation of charge compensated sites in CaF2: Er3+,” J. Chem. Phys. 63(5), 2074–2085 (1975). [CrossRef]  

41. S. Normani, A. Braud, R. Soulard, J. L. Doualan, A. Benayad, V. Menard, G. Brasse, R. Moncorgé, J. P. Goossens, and P. Camy, “Site selective analysis of Nd3+–Lu3+ codoped CaF2 laser crystals,” CrystEngComm 18(47), 9016–9025 (2016). [CrossRef]  

42. C. Bernerd, B. D. C. Fernandes, E. Boursier, J. Luce, C. Maunier, A. Braud, C. Meroni, P. Camy, and S. Montant, “Thermal lens in diode-pumped square-section rods of Nd-doped phosphate glass and 0.5%Nd:5%Lu:CaF2 crystal: simulations and experiments,” Appl. Opt. 59(31), 9905–9911 (2020). [CrossRef]  

43. N. P. Barnes and B. M. Walsh, “Quantum efficiency measurements of Nd: YAG, Yb: YAG, and Tm: YAG,” Adv. Solid-State Lasers 68, 284–287 (2002).

44. A. A. Demidovich, A. P. Shkadarevich, M. B. Danailov, P. Apai, T. Gasmi, V. P. Gribkovskii, A. N. Kuzmin, G. I. Ryabtsev, and L. E. Batay, “Comparison of cw laser performance of Nd: KGW, Nd: YAG, Nd: BEL, and Nd: YVO4 under laser diode pumping,” Appl. Phys. B 67(1), 11–15 (1998). [CrossRef]  

45. N. P. Barnes, M. E. Storm, P. L. Cross, and M. W. Skolaut, “Efficiency of Nd laser materials with laser diode pumping,” IEEE J. Quantum Electron. 26(3), 558–569 (1990). [CrossRef]  

46. L. G. DeShazer and R. S. Pierre, “Novel neodymium hosts,” in Conference on Lasers and Electro-Optics, (Optical Society of America, 1992), CTuG1.

47. D. G. Matthews, J. R. Boon, R. S. Conroy, and B. D. Sinclair, “A comparative study of diode pumped microchip laser materials: Nd-doped YVO4, YOS, SFAP and SVAP,” J. Mod. Opt. 43(5), 1079–1087 (1996). [CrossRef]  

48. H. He, X. S. Liu, Y. H. Song, C. C. Wang, M. Z. Cao, A. R. Yan, and Z. Y. Wang, “LD end-pumped Nd:YVO4 high energy high beam quality 1064 nm picosecond laser with a semiconductor saturable absorber mirror,” Optik 175, 172–176 (2018). [CrossRef]  

49. Y. Sato and T. Taira, “Comparative study on the spectroscopic properties of Nd:GdVO4 and Nd:YVO4 with hybrid process,” IEEE J. Sel. Top. Quantum Electron. 11(3), 613–620 (2005). [CrossRef]  

50. T. de Almeida Vieira, F. M. Prado, and N. U. Wetter, “Near Quantum Limited Slope Efficiency Nd:YLFF4 Laser,” Adv. Solid-State Lasers, AW3A.6 (2021).

51. N. Wetter, “Neodymium doped lithium yttrium fluoride (Nd:YLiF4) lasers,” in Handbook of Solid-State Lasers (Elsevier, 2013), pp. 323–340.

52. W. Koechner, Solid-state Laser Engineering (Springer, 2013), Vol. 1.

53. Y. F. Chen and Y. P. Lan, “Comparison between c-cut and a-cut Nd:YVO4 lasers passively Q-switched with a Cr4+:YAG saturable absorber,” Appl. Phys. B 74(4-5), 415–418 (2002). [CrossRef]  

54. Z. B. Pan, H. J. Cong, H. H. Yu, L. Tian, H. Yuan, H. Q. Cai, H. J. Zhang, H. Huang, J. Y. Wang, and Q. Wang, “Growth, thermal properties and laser operation of Nd:Ca3La2(BO3)4: A new disordered laser crystal,” Opt. Express 21(5), 6091–6100 (2013). [CrossRef]  

55. T. Suzuki, H. Kawai, H. Nasu, M. Hughes, O. Yasutake, S. Mizuno, H. Ito, and K. Hasegawa, “Quantum efficiency of Nd3+-doped glasses under sunlight excitation,” Opt. Mater. 33(12), 1952–1957 (2011). [CrossRef]  

56. K. Naito, M. Yamanaka, M. Nakatsuka, T. Kanabe, K. Mima, C. Yamanaka, and S. N. S. Nakai, “Conceptual design studies of a laser diode pumped solid state laser system for the laser fusion reactor driver,” Jpn. J. Appl. Phys. 31(2R), 259 (1992). [CrossRef]  

57. A. Agnesi, A. Greborio, F. Pirzio, E. Ugolotti, G. Reali, S. Y. Choi, F. Rotermund, U. Griebner, and V. Petrov, “Femtosecond Nd: glass lasers pumped by single-mode laser diodes and mode locked with carbon nanotube or semiconductor saturable absorber mirrors,” IEEE J. Sel. Top. Quantum Electron. 18(1), 74–80 (2012). [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 (8)

Fig. 1.
Fig. 1. The thermodynamic stable clusters in La3+-, Nd3+-, Gd3+-, Y3+-, Lu3+-doped CaF2 crystals, these centers are divided into three groups, “1”, “2” and “3” represent the cubic monomers, cubic sublattice clusters and square antiprism structure clusters, respectively.
Fig. 2.
Fig. 2. The number of rare earth ions dependent formation energy of the thermodynamic stable centers.
Fig. 3.
Fig. 3. Thermodynamic stable centers of the mixed [Nd3+-La3+] and [Nd3+-Y3+] clusters in CaF2 crystal.
Fig. 4.
Fig. 4. Number of R3+ ions dependent formation energy of the mixed [Nd3+-R3+] clusters in CaF2 crystal, R = La (a) and Y (b), respectively, the m and n represent number of Nd3+ and R3+ atoms in the clusters, respectively.
Fig. 5.
Fig. 5. pDOS of the lanthanide and interstitial fluorine in 11|0|0|11 center of CaF2. (a) The states of La3+ (4f), Nd3+ (4f) and Y3+ (4d); (b) the 2p states of the Fi-.
Fig. 6.
Fig. 6. Comparison of area normalized absorption (a), intensity normalized emission spectra (b) and (c) decay curves of Nd3+,R3+:CaF2 crystal.
Fig. 7.
Fig. 7. The codopant concentrations dependent photoluminescence parameters of Nd3+,R3+:CaF2 crystals, (a) the quantum efficiencies, (b) the fluorescence lifetimes, (c) the peak emission cross sections.
Fig. 8.
Fig. 8. The local structures dependent peak emission cross sections and lifetimes in 0.5%Nd3+,8%R3+:CaF2 crystals.

Tables (4)

Tables Icon

Table 1. Formation energy of the thermodynamic stable rare earth clusters in CaF2 crystal.

Tables Icon

Table 2. The formation energy (FE), averaged bond length (ABL), coordination number (CN) and the ionic sublattice (SBL) around Nd3+ in [Nd3+-La3+] centers, the SAP means square antiprism sublattice.

Tables Icon

Table 3. The formation energy (FE), averaged bond length (ABL), coordination number (CN) and the ionic sublattice (SBL) around Nd3+ in [Nd3+-Y3+] centers, the SAP means square antiprism sublattice.

Tables Icon

Table 4. Comparison of the photoluminescence parameters and thermal conductivities at room temperature of the neodymium-doped laser materials.

Equations (4)

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

Δ E = ( E t o t + E 0 ) i 1 E 1 i 2 E 2 ( s v ) E 3 { [ i 1 + i 2 + ( s v ) ] [ i 1 + i 2 + ( s v ) ] 2 } E corr
E corr = ( 1 + g ) q 2 α 2 ε L
τ e m = t = 0 I ( t ) t d t t = 0 I ( t ) d t
σ e m = λ p e a k 4 8 π c n 2 β 2 η τ e m Δ λ e m
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.