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

Tailoring the local lattice distortion of Nd3+ by codoping of Y3+ through first principles calculation for tuning the spectroscopic properties

Open Access Open Access

Abstract

Trivalent rare earth ions have rich and unique energy levels arising from 4f inner shell configurations, which can be widely used in lasers, display, and bioimaging, etc. As the f-f transition of the rare earth ions is parity-forbidden, it is necessary to dope the active ions in the host with an appropriate coordination structure to relax the forbidden. In this work, a first principles calculation based on the density functional theory was applied to investigate the local structure symmetries of Nd3+ ions in Nd3+,Y3+:SrF2 crystal. The computational results show that the clusters of [mNd3+-nY3+] (m + n = 2, 3, 4, 5 and 6) would be formed when Y3+ is codoped in Nd3+:SrF2 crystal. The formation energy of the [mNd3+-nY3+] cluster decreases when the value of m + n increases, or the value of n increases with m + n fixed, then the lattice structure becomes more stable. Furthermore, the first coordination shell of Nd3+ was cubic when the n ≤ 1, and it would transform to be the lower symmetric square antiprism structure with n ≥ 2. The forbidden of the electric dipole transition was thus relaxed due to breaking of the symmetry. The experimental results show the absorption cross section of Nd3+ was increased from 0.45 ×10−20 cm2 to 5.2×10−20 cm2, and the emission intensity was also enhanced by about 20 times through codoping Y3+ ions, which agreed well with the calculated results. It suggests that the tailoring of the local lattice distortion of the active ions may open interesting possibilities in the design of rare earth doped materials.

© 2019 Optical Society of America under the terms of the OSA Open Access Publishing Agreement

1. Introduction

Since the first working laser based on a ruby crystal was built in 1960 [1], the development of laser crystal materials has undergone for over half a century. Nowadays, as in the past, the bulk laser crystals play a crucial role in laser technologies and engineering, although the investigation and development of new gain materials, such as fiber, waveguides and transparent ceramics [2], grow very fast. Generally, the laser crystals consist of active ions and host materials. There are hundreds of laser crystals being formed by a combination of the active ions of rare-earth ions or transition metal ions, with oxide and fluoride host crystals [3]. However, as regards practical lasers their number is very limited, and the development of new laser crystals in the manner is to some degree unsustainable.

New laser crystals are highly desirable in novel laser technologies and applications [4]. Several new strategies, including cationic/anionic substitution [5], chemical unit cosubstitution, cation nano-segregation and local structure manipulation were tried and extensively applied on the photoluminescence properties tuning of phosphor materials [68]. By manipulating local coordination environment of the active ions, realization of tuning spectral properties was highly expected. The concept of local structure manipulation provides a possible methodology, and has been tentatively tried in tuning spectral properties and laser performances of rare-earth-doped fluorite-structure crystals [917].

Due to the charge compensation effects, required to accommodate a trivalent ion in the divalent cationic sites, the doped sites of rare-earth ions become inequivalent, also trivalent ion clusters were easily formed in fluorite crystals [1819]. The clustering processes were severely rare-earth concentration and host lattice dependent [2021]. Signals of 18 kinds of centers were observed in the crystal [22]. Therefore, rare-earth doped fluorite-like crystals have very broad absorption and emission band full width at half maximum (FWHM) caused by these centers. Broad emission made the crystals particularly appealing for generation of ultrashort pulses. Additionally, due to the abundant 4f, 5d levels, the energy relaxations will bring benefits for the emissions and laser performances at 1.8 µm of Tm3+ and 2.8 µm of Er3+ [1415]. However, luminescence at 1 µm of Nd3+ has been deadly quenched by the incoherent dipole-dipole energy transfer process [23]. Kaminskii et al. firstly introduced Y3+, La3+, Gd3+ and Sc3+ to break the quenching clusters and ameliorate the quantum yields in 1960s [2425], and they were called buffer ions [26]. Furthermore, Catlow, Cheetham et al. have devoted themselves on the clusters study by both modeling and experiment methods. Centers of monomers, dimers, trimers, tetramers and hexamers were simulated in CaF2, SrF2 and BaF2, and the stabilities varied with the dopants and matrix crystals [1819,27]. Local environment of the centers altered with size of the rare-earth ion, and signals of dimer and hexamer clusters were observed by EXAFS [28]. Two kinds of monomers were identified and their relative concentrations in Gd3+ doped CaF2, SrF2 and BaF2 were studied by the correlated EPR and ITC technique [29]. The neutron diffraction results show that there are vacancies in normal fluorine lattice sites, and two types of interstitials Fi were detected. Variations in relative number of these defects intimately connected with local structures and configurations of the clusters [30].

On the base of these achievements, from beginning of 21st century, a lot of important progresses have been obtained in rare-earth doped fluorite crystal. For example, in Yb3+-doped calcium fluoride (CaF2), the spectral properties were modulated by Na+ in large scale, and 3-mJ femtosecond pulses were realized [910]. The FWHM of the emission band of Nd3+ in CaF2 crystal was manipulated by Y3+ to be 31 nm, and mode-locked laser pulses of 103 fs were obtained [11]. The FWHM of the band is even comparable with that of Nd3+-doped glasses [31]. By codoping Gd3+ in Pr3+:CaF2 the photoluminescence was highly improved and red lasing was firstly achieved [12]. With addition of Na+ the lasing wavelength of Tb3+:CaF2 was continuously tuned from 539.5 to 550.0 nm in green range [13]. The laser tuning range of more than 190 nm was firstly demonstrated in Tm3+:CaF2 when codoped with Y3+ [14]. Highly efficient dual-wavelength lasing at 2.8 µm was performed in Er3+-doped strontium fluoride (SrF2) [15], and the doping concentration was an order of magnitude lower than that of Er3+:YAG crystal [16]. For Nd3+:SrF2, the emission band FWHM was modulated to be 20 nm when codoped with Y3+, and pulse of 97 fs was achieved [17]. The spectral properties and laser performances were realized to be tuned in large scope by manipulating local environment of the active ions in fluorite crystals. However, pictures about local structure of the active ions manipulated by Sc3+, Y3+, Ce3+, La3+, Gd3+, Lu3+, O2-, Na+ and K+, as well as the other trivalent rare-earth ions, were unknown after nearly six decades.

In this article, local structures of the active Nd3+ tailored by Y3+ were studied by first principles calculation. It was found that the cubic coordination surroundings of Nd3+ were transformed to be distorted lower symmetric square antiprism structures. The more relaxed parity forbiddance made the f-f transition rate and near-infrared emissions be highly enhanced. Moreover, low temperature TRES suggests that a multisite crystal was changed to be a quasi-single center system. From this perspective, the methodology of local structure manipulation could offer new opportunity to tune spectral properties and performances of rare-earth doped fluorite materials.

2. Materials and experiments

Crystals of 0.5 at.% Nd3+, x at.% Y3+:SrF2 (x = 0, 1.5, 2, 4.5, 5 and 10 denoted as A, B, C, D, E and F, respectively), y at.% Nd3+:SrF2 (y = 0.09, 0.4, 0.6, 0.85 and 1.65 as A’, B’, D’, E’ and F’, respectively) and y at.% Nd3+,5 at.% Y3+:SrF2 (y = 0.09, 0.4, 0.6, 0.85 and 1.65 as A’’, B’’, D’’, E’’ and F’’, respectively) were synthesized by temperature gradient technique method. The concentration of Y3+ was 50 times higher to ensure that all Nd3+ centers were changed to be [Nd3+-Y3+] clusters. Raw materials of the fluorites with purity of 99.99% were finely mixed and loaded in the graphite crucible. The melting temperature is about 1693 K, the growth with rate of −1.5 K/hour was completed in the vacuum of 10−3 Pa, and finally the crystals were obtained. The part near the seed was cut and prepared in the same dimensions for all the measurement.

In the structure relaxations, the plane-wave basis set and the Perdew-Burke-Ernzerhof exchange correlation potential have been used as implemented in the VASP code [3236]. The projector augmented wave (PAW) potential was used to describe the electron-ion interactions [3738]. In addition, the spin polarization was included in all calculations. A 550 eV cut-off in the plane-wave expansion and a 1×1×1 Gamma k-grid were used to ensure that the relaxation has an accuracy of 10−5 eV. The internal coordinates of the supercell (of size 2×2×2) were optimized until forces on individual atoms reached 0.01 eVÅ−1 to obtain sufficient accuracy. Formation energy of the cluster (ΔE) was obtained based on formula (1).

$$\begin{array}{c} \Delta E = ({E_{tot}} + {E_0}) - m \cdot {E_1} - n \cdot {E_2} - w \cdot {E_3}\\ - [m + n + w - {(m + n - w)^2}] \cdot {E_{corr}} \end{array}$$
Where Etot represents the total relaxed energy of the supercells including [mNd3+-nY3+] centers; m, n and w respectively the number of Nd3+, Y3+ and interstitial fluorine within a cluster; E0 the relaxed energy of pure SrF2 crystal; E1, E2 and E3 respectively the relaxed energy of the 2×2×2 supercells including [1Nd3+−0Fi], [1Y3+−0Fi] and [0Nd3+−1Fi] center; Ecorr is the potential alignment and image charge corrections calculated by Eq. (2) [3941].
$${E_{corr}} = (1 + g)\frac{{{q^2}\alpha }}{{2\varepsilon L}}$$
Where g is the scaling factor, for face-centered cubic structure the value of −0.34 was adopted [41], q is the net charge, α the Madelung constant of 5.0388 [42], ɛ and L respectively the static electric constant and supercell dimensions of SrF2 crystal, ɛ equals to 6.47 at 300 K [43], 11.6 Å was chosen for L. For a 2×2×2 supercell of SrF2 crystal with a net charge (q = 1) the value was calculated to be 0.319 eV. The projected density of state (pDOS) calculations were performed based on the optimized structure with a 4×4×4 Gamma k-grid. The Dudarev’s LSDA + U method was considered for the on-site Coulomb interaction [44]. In the work U = 1 and U = 6 were chosen for yttrium and neodymium, respectively.

Room temperature absorption spectra were recorded by a Jasco V-570 UV/VIS/NIR spectro-photometer and the photoluminescence spectra were tested with a FLS980 time-resolved fluorimeter grating blazed at 1200 nm and detected by thermoelectric (TE) cooled InGaAs detector. The excitation was a xenon lamp at 796 ± 5 nm. The lifetime was measured using the fluorimeter and Tektronix TDS 3052 oscilloscope following a flash lamp excitation at 796 ± 5 nm. Time resolved emission spectra (TRES) was realized by testing the decay curves using the same fluorimeter following a flash lamp excitation at 796 ± 5 nm. The detected wavelength ranged from 1020 to 1120 nm with steps of 0.5 nm and integrating time of 20 s. The samples were placed into a Linkam THMS600E (UK) continuous-flow liquid nitrogen cryostat. The temperature was controlled at 78 K by the Linksys32 system.

3. Results

3.1 Rare earth ion clusters simulation

According to the references of [18] and [19], centers of [mNd3+] and [nY3+] from monomer to hexamer (m = 1∼6, n = 1∼6) were modelled. However, no related content about [Nd3+-Y3+] cluster was available, and all possible [mNd3+-nY3+] supercells were tentatively constructed and then relaxed using the VASP code [3235]. The screening, based on the thermodynamic stability computation, was performed. It can be seen in Figs. 12 and Table 1, Nd3+ ions, as well as Y3+, are easily clustering in SrF2 crystal. The centers were divided into three groups, the first is monomer centers, and the others are clusters but with different local sublattice, cubic and square antiprism, respectively. In Nd3+:SrF2 crystal, the stability of [1Nd3+−1Fi] centers with interstitial Fi at the nearest and next nearest site is the same. While for Y3+:SrF2 the next nearest site center is more stable, agreed well with the reported results [18]. There are dimer, trimer and tetramer cubic sublattice clusters in Nd3+-doped SrF2, and only dimer centers existed in Y3+-doped crystal, as demonstrated in Figs. 12. The difference should be ascribed to the competition of Nd3+-Fi or Y3+-Fi attraction with Fi−Flattice (matrix lattice) repulsion, which will be discussed further.

 figure: Fig. 1.

Fig. 1. Thermodynamic stable Nd3+ centers in SrF2 crystals: (a) first group monomer centers, (b) second group clusters with cubic sublattice local environment and (c) third group clusters and the local is square antiprism. The bracket suffixed with different numbers means the same centers with different configurations.

Download Full Size | PDF

 figure: Fig. 2.

Fig. 2. Thermodynamic stable Y3+ centers in SrF2 crystals: (a) first group monomer centers, (b) second group cubic sublattice clusters and (c) third group square antiprism clusters. The bracket suffixed with different numbers means the same centers with different configurations.

Download Full Size | PDF

Tables Icon

Table 1. Simulated formation energy of Nd3+ or Y3+ centers in SrF2 crystals. The bracket suffixed with different numbers means the same centers with different configurations, “ / ” denotes that the center was not stable.

[Nd3+-Y3+] clusters were also simulated, as presented in Fig. 3 and Table 2. Nd3+ and Y3+ also tend to aggregate into clusters of [mNd3+-nY3+]. Due to the strong repulsion of Fi−Flattice, nearly all [Nd3+] centers are surprisingly changed to be the square antiprism sublattice clusters. It is noted that the cubic and square antiprism sublattice coexisted in [2Nd3+−1Y3+−4Fi] center, presenting clustering characters of both Nd3+ and Y3+. When more than one Y3+ ions (n ≥ 1) joined in a cluster the cubic sublattice would be transformed to be square antiprism to relieve the inner strain, as shown in Fig. 3(b).

 figure: Fig. 3.

Fig. 3. Thermodynamic stable [Nd3+-Y3+] centers in SrF2 crystal: (a) second group cubic sublattice clusters, (b) third group square antiprism clusters. The bracket suffixed with different numbers means the same centers with different configurations.

Download Full Size | PDF

Tables Icon

Table 2. Simulated formation energy of [Nd3+-Y3+] clusters in SrF2 crystal. The bracket suffixed with different numbers means the same centers with different configurations.

Formation energy of the clusters were plotted and presented in Fig. 4(a). The energy of [Nd3+-Y3+] clusters decreased linearly with increasing number of Y3+ ions introduced in the center. The slope of [Nd3+-Y3+] clusters was nearly identical with that of pure Y3+ centers. As shown in Fig. 4(b), formation energy will be reduced by 1.113 eV with one more Nd3+ added in the cluster, and it is 1.400 eV for Y3+. Furthermore, the formation energy of [mNd3+-nY3+] clusters with m = 1 and n respectively equals to 1, 4 and 5 in Fig. 4(a), is even lower than that of pure Y3+ centers, which means that Y3+ is a very potential and efficient regulator to recombine with Nd3+ ions. Simultaneously, the local surroundings were modulated to be more distorted, as shown in Fig. 4(c). [3Y3+−4Fi] and [4Y3+−4Fi] centers are more stable than that of [1Nd3+−2Y3+−4Fi] and [1Nd3+−3Y3+−4Fi] clusters, respectively. Some portion of the introduced Y3+ would aggregate into pure Y3+ centers rather than joined together with Nd3+.

 figure: Fig. 4.

Fig. 4. (a) Formation energy of Nd3+ and/or Y3+ clusters versus the number of Y3+ ions within a cluster. Hollow symbol with the same solid shape represents the same clusters with different configurations. (b) Number of dopant cations dependent formation energy of Nd3+ or Y3+ centers. (c) Local distortion of Nd3+ in [Nd3+−Y3+] polyhedron, the arrow with directions to Nd3+ and F means shortened and elongated bond length, respectively, when compared with the pure Nd3+ centers.

Download Full Size | PDF

Thus clear pictures about local structure of [Nd3+] centers tailored by Y3+ were obtained. Besides Y3+, how about other ions, are they playing the same role. For the answers, it should be known firstly why the local environment of Nd3+ could be tailored by Y3+. The projected DOS was computed, as demonstrated in Figs. 5(a)–(b) of [1Y3+−1Fi] center. The partial bond connectivity between Y3+ and interstitial Fi was established when Fi at the nearest site, and no such bonding occurred while at the next nearest site which means that only coulomb interaction existed in Fig. 5(b). Due to Y3+−Fi attraction, distance of Fi and Flattice in Fig. 5(a) was shortened. Most of the nonbonding electrons were localized on interstitial Fi rather than extended to the vacant lattice cation orbitals or the dopant cations. This made the repulsion force of Fi−Flattice larger than that of Y3+−Fi attraction when interstitial Fi at the nearest site. Formation energy of [1Y3+−1Fi] center with Fi at the nearest site was higher about 0.2 eV than that at the next nearest site in virtue of large inner strain. For Y3+ clusters with several interstitials Fi and substitutional impurities, due to the strong repulsion of Fi−Flattice and the relatively weak Y3+−Fi interaction, it was sufficient to drive the cubic local sublattice to be square antiprism in Fig. 3(b). As can be seen in Fig. 9(a) (Appendix), p state of Fi was completely bonded with the dopant cation in square antiprism Y3+ clusters.

 figure: Fig. 5.

Fig. 5. Projected DOS on local coordination structure of the cluster. The p state of fluorine in [1Y3+−1Fi] and [1Nd3+−1Fi] centers with interstitial Fi at the nearest site (a) and (c), and the next nearest site (b) and (d), respectively. The insert represents p, d and f states of the corresponding rare-earth ions.

Download Full Size | PDF

However, it is a little different between Nd3+ and Y3+ centers. As can be seen in Figs. 5(c)–(d) and Fig. 10 (Appendix), p state of interstitial Fi at the nearest site was hybridized with d, f and p states of Nd3+, as well as d state of the nearest strontium atom. The electrons of Fi would be delocalized to levels on Nd3+ and the conduction band. As a result the interaction of Nd3+ and Fi was strengthened. At the same time, the lowered charge density made Fi−Flattice repulsion force weakened and then the inner strain relieved. Catlow et al. reported that covalent bonding between the interstitials Fi could possibly be occurred [45]. The projected density of state results suggest that if electron density of Fi was largely removed, attraction interaction between Fi−Fi would make the center more stable. It was thus stabilized [1Nd3+−1Fi] center with Fi at the nearest site, as well as the cubic sublattice trimer clusters. The square antiprism sublattice tetramer, pentamer and hexamer centers also existed, and full bonding was observed in Fig. 9(b) (Appendix).

The local environment tailoring originates from the big difference of bonding characters of Nd3+−Fi and Y3+−Fi, in other words, by choosing appropriate ions with desired bonding characters, the target of local environment and spectral properties manipulation could be realized.

3.2 Tailored photoluminescence of the crystal

Based on the results, a strategy of local structure manipulation was proposed and illustrated in Fig. 6. The cubic first coordination shell of Nd3+ was tailored by the regulators to be lower symmetric square antiprism sublattice while the matrix keeps in fluorite structure. As shown in Fig. 11 (Appendix), the XRD patterns show that all the crystals with high crystallinity remain in fluorite-like structure, and diffraction of (111) as well as the other planes inclined and shift to higher angles, which indicates that the interplanar distance and the size of the cells were shrinked. The process of local structure transformation from cubic to square antiprism cancelled out the expansion induced by the interstitials Fi in the lattice, which contributed to the contracted cells. The lowered symmetric square antiprism sublattice tailored by Y3+ made the electric dipole transitions of f-f be therefore greatly enhanced, and good laser performance was expected. The demands for new ultrashort and tunable laser materials were highly expected to be realized in the trivalent lanthanide-doped fluorite crystals, rather than changing and trying different matrix crystals tediously as it did before, and the assessment about the dopants induced negative influence that usually diminishes the emission quantum yields should also be considered.

 figure: Fig. 6.

Fig. 6. Strategy of local structure manipulation. With no loss of crystal structure integrity, the cubic first coordination shell of active Nd3+ (blue core shell) would be manipulated to be distorted lower symmetric square antiprism sublattice when more than one Y3+ ion (green core shell) was introduced.

Download Full Size | PDF

The peak absorption cross sections calculated by absorption coefficient and real concentration of Nd3+ in the crystal measured by inductively coupled plasma-optical emission spectroscopy (ICP-OES) of Nd3+:SrF2 [46] increased with increasing concentration of Nd3+, as well as product of the cross section and FWHM, as shown in Fig. 7(a) and Fig. 12 (Appendix). The big clusters are more thermal-dynamically stable in Fig. 4(a), which means that the clustering process was dopant concentration dependent. With increasing concentration the high symmetric monomers were transformed to be the lower symmetric clusters in Fig. 1, and the cross sections were thus enhanced gradually. With addition of 5% Y3+, the cross section, as well as the product was greatly enhanced, and continuously increased with decreasing concentration of Nd3+. The enhanced peak absorption cross section and the area of the absorption band were in well consistent with the calculated results that the local environment was tailored to be the lower symmetric square antiprism structure with addition of Y3+. As illustrated in Figs. 4(c) and 6, the seriously distorted local structure of Nd3+ contributed to the significantly enhanced emission in samples with 5% Y3+. It can be seen in Fig. 7(b), the intensity was enhanced by about 20 times. Taking the doping concentration of Nd3+ into consideration, the intensity follows nearly the same pattern as that of absorption cross section in Fig. 7(a).

 figure: Fig. 7.

Fig. 7. Nominal molar concentration of Nd3+ dependent peak absorption cross section (a) and normalized PL intensity at 1057 nm (b) of y at.% Nd3+:SrF2 and y at.% Nd3+,5 at.% Y3+:SrF2. (c) Y3+ concentration dependent peak absorption cross section of 0.5 at.% Nd3+, x at.% Y3+:SrF2. (d) PL spectra of 0.5 at.% Nd3+, x at.% Y3+:SrF2, the intensity of sample A was multiplied by 1.5 for clear view. The insert is Y3+ concentration dependent normalized emission intensity at 1057 nm. Instrumental: SBW = 8 cm−1. The samples in same dimensions were measured under the same conditions, and the emission intensity could therefore be compared with each other.

Download Full Size | PDF

One more group of Nd3+,Y3+:SrF2 crystal was prepared to ensure the repeatability. As can be seen in Figs. 7(c)–(d) and Fig. 13 (Appendix), the absorption cross section and emission were also greatly improved with addition of Y3+ in 0.5% Nd3+:SrF2. The emission is very weak in Nd3+ alone doped crystal. Notably, when incorporated with Y3+, the emission peaks at 1036, 1044, 1075 and 1084 nm of Fig. 7(d), associated with long lifetime component in Fig. 14 (Appendix) [47], were gradually eliminated and intensity boosted up at 1057 nm. The long and short components of lifetimes, respectively connected with Nd3+ monomers (Fig. 1(a)) and clusters (Figs. 1(b)–(c)) [48], were finally merged to be one value. At the same time, due to the decreased inhomogeneous broadening effect caused by the decreased number of Nd3+ sites, the FWHM of emission decreased exponentially from 28 to 20.5 nm, which is competent for sub-100 fs pulses generation [17], as shown in Fig. 15 (Appendix). That agreed qualitatively with the changing of simulated [Nd3+−Y3+] clusters in Figs. 3 and 4(a). It is noted when the concentration of Y3+ is above 5 at.% the values tend to be constant. It is probably because of the saturated emission of Nd3+ in square antiprism lattice of Fig. 3(b).

The absolute quantum yields were also measured and presented in Fig. 16 (Appendix). The efficiency increased exponentially with concentration of Y3+, a value of 11.3% of 0.5% Nd3+:SrF2 was improved to be 88.2% due to the combination of [Nd3+−Y3+] clusters in Fig. 3. The quantum yields (η) and fluorescence lifetime (τexp) were utilized to calculate the radiative lifetime (τrad) based on η= τexprad. For the curves of C and D crystal respectively with 2 at.% and 4.5 at.% Y3+, the deviation from single exponential profile was relatively small, and the mean lifetime was adopted [49]. Thus, the two samples together with E and F were included in the calculation. As shown in Fig. 17 (Appendix), the radiative lifetime decreased exponentially with addition of Y3+. As can be seen in Figs. 3(b), 4(c) and 6, the transformation from cubic to distorted square antiprism sublattice of Nd3+ made the parity forbiddance be greatly relaxed. The f-f transition rate was therefore increased exponentially. Similarly, the peak emission cross section, evaluated by Füchtbauer-Ladenburg equation [50], was also tunable, as shown in Fig. 18 (Appendix).

3.3 Low temperature TRES analyses

To better understand the clustering process, A, C and F crystal were chosen for TRES measurement in liquid nitrogen, 3-dimensions (3D) image and 2D TRES were presented in Fig. 8 and Fig. 19 (Appendix), respectively. Due to the inhomogeneous broadening caused by the Nd3+ clusters covered from monomers (Fig. 1(a)) to clusters (Figs. 1(b)–(c)), the photoluminescence was observed to be very broad in 0.5% Nd3+:SrF2 crystal. With addition of Y3+ from 2 at.% to 10 at.% signals of the monomers were gradually disappeared, and square antiprism [Nd3+−Y3+] clusters finally dominated in the crystal. The emission peaks were manipulated to be nearly identical with each other, which means a multisite crystal was changed to be a quasi single center system by codoping the lattice regulator ions of Y3+. Nd3+ centers would be recombined with Y3+ and transformed to be the third group of lower symmetric square antiprism clusters. The nearly identical first coordination shell of the active Nd3+ was then obtained. However, various configurations of the cluster made the splitting levels slightly different and thus the FWHM inhomogeneously broadened. The transitions were severely overlapped, it is difficult to split them even at liquid helium temperature. The aggregating processes show that the transformation was Y3+ concentration dependent. Although high order clusters are more stable in thermodynamics, relatively high concentration doping of Y3+ or Nd3+ should be essential.

 figure: Fig. 8.

Fig. 8. Low temperature TRES on clustering process. 3D image of the TRES of 4F3/24I11/2 transition of A (a), C (b), and F (c) samples. The insert represents the viewing along z-axis.

Download Full Size | PDF

4. Discussion

Based on the aforementioned results, the clusters of [Nd3+−Y3+] are more stable than that of corresponding [Nd3+−Nd3+] clusters. It was confirmed that the fluorescence quenching could be eliminated with addition of Y3+ ions [2426]. More interestingly, the cubic local structure was surprisingly driven to be the distorted square antiprism structure and the “phase transition” occurred in the sub-nanometer particles (clusters), and a multisite crystal was manipulated to be a quasi-single center system, which suggested that Y3+ contributed as local structure regulator ions. The local structure transformation depends on the coupled interactions induced by the delocalization of charge density of interstitials Fi in the clusters. For different rare-earth ions and host fluorite materials, the Fi−Flattice, Fi−Fi interactions and the bonding of Fi with the dopant cations are different. The varied local environments are thus expected in rare-earth doped fluorites.

Furthermore, the geometry optimization in the work was all simulated in equilibrium state. It suggests, that local environment of the dopant cations would also be influenced by preparation processes and heat treatment routines [51], as well as the dopants concentrations [2021]. It is unknown but will be very interesting about the local structures and configurations in the nonequilibrium states, such as the preparation temperatures, pressures, electric and magnetic fields. The local structures of the clusters would be evolved dynamically with such force fields. Direct observation and characterization of the clusters is another interesting topic, and the work is on the way. Nevertheless, designing of laser crystal with tunable spectral properties could be realized by tailoring the rare earth ion clusters in fluorites.

5. Conclusion

In summary, an efficient local structure manipulation strategy was proposed by tailoring local structure distortion of the active ions in fluoride crystal. By virtue of strong repulsion of Fi−Flattice interaction, the first shell of Nd3+ with cubic sublattice was manipulated to be lower symmetric square antiprism structure in [Nd3+−Y3+] cluster with no loss of crystal structure. The distorted low symmetric quasi single center system contributed to the greatly enhanced emissions. Our results suggest that local structure manipulation approach could be useful in design of new laser materials with controlled optical spectra properties.

Appendix

Characterization

XRD testing

X-ray diffraction (XRD) was measured by Rigaku Ultima-IV diffractometer (Japan). Cu Kα radiations and scan step of 0.02 ° were used in the angle region of 20 ° to 70 ° on 2θ scale, and signals caused by Kα2 radiation were striped from the data.

Absolute quantum yields measurement

The absolute quantum yields were tested with a FLS980 time-resolved fluorimeter grating blazed at 1200 nm and detected by thermoelectric (TE) cooled InGaAs detector. The excitation was at 796 ± 15 nm. A barium sulfate-coated integrating sphere (Edinburgh) was applied as the sample chamber. The entry and output gates of the sphere located in 90 ° geometry from each other in the plane of the fluorimeter. Both reference and sample emissions were collected in the region of 765 to 1450 nm. The yields were thus obtained.

 figure: Fig. 9.

Fig. 9. Projected DOS of p state of interstitial Fi in the square antiprism lattice of (a) Y3+ clusters and (b) Nd3+ centers.

Download Full Size | PDF

 figure: Fig. 10.

Fig. 10. Projected DOS of d state of Sr atom nearest to [1Nd3+−1Fi] center with Fi at the nearest site.

Download Full Size | PDF

 figure: Fig. 11.

Fig. 11. XRD patterns of the crystal.

Download Full Size | PDF

 figure: Fig. 12.

Fig. 12. Nominal Nd3+ concentration dependent product of peak absorption cross section (σabs) and FWHM (λabs) of 4I9/24F5/2 + 2H9/2 transition.

Download Full Size | PDF

 figure: Fig. 13.

Fig. 13. Absorption cross section of the crystal, all peaks were attributed to the absorption of Nd3+ ions, the strongest band is around 800 nm corresponding to 4I9/24F5/2 + 2H9/2.

Download Full Size | PDF

 figure: Fig. 14.

Fig. 14. Logarithmic decaying curves of the sample, for A, B, C and D crystals the lifetime consists of a long and short component, while for E and F the decaying could be well fitted by single exponential expression.

Download Full Size | PDF

 figure: Fig. 15.

Fig. 15. Nominal Y3+ concentration dependent FWHM of 4F3/24I11/2 transition.

Download Full Size | PDF

 figure: Fig. 16.

Fig. 16. Nominal Y3+ concentration dependent absolute quantum yields of 0.5 at.% Nd3+, x at.% Y3+:SrF2 crystal.

Download Full Size | PDF

 figure: Fig. 17.

Fig. 17. Nominal Y3+ concentration dependent fluorescence and radiative lifetime of 0.5 at.% Nd3+, x at.% Y3+:SrF2.

Download Full Size | PDF

 figure: Fig. 18.

Fig. 18. Nominal Y3+ concentration dependent peak emission cross section of 0.5 at.% Nd3+, x at.% Y3+:SrF2 crystal.

Download Full Size | PDF

 figure: Fig. 19.

Fig. 19. 2D-TRES was ten sliced, integrated and normalized spectra of A (a), C (b), and F (c) crystal.

Download Full Size | PDF

Funding

National Natural Science Foundation of China (61635012, 61905289); National Key Research and Development Program of China (2016YFB0402101); Strategic Priority Program of the Chinese Academy of Sciences (XDB16030000).

References

1. T. H. Maiman, “Stimulated optical radiation in ruby,” Nature 187(4736), 493–494 (1960). [CrossRef]  

2. A. A. Kaminskii, “Laser crystals and ceramics: recent advances,” Laser Photonics Rev. 1(2), 93–177 (2007). [CrossRef]  

3. A. A. Kaminskii, “Modern developments in the physics of crystalline laser materials,” Phys. Status Solidi A 200(2), 215–296 (2003). [CrossRef]  

4. P. Norvig, D. A. Relman, D. B. Goldstein, D. M. Kammen, D. R. Weinberger, L. C. Aiello, G. Church, J. L. Hennessy, J. Sachs, and A. Burrows, “2020 Visions,” Nature 463(7277), 26–32 (2010). [CrossRef]  

5. W. T. Chen, H. S. Sheu, R. S. Liu, and J. P. Attfield, “Cation-size-mismatch tuning of photoluminescence in oxynitride phosphors,” J. Am. Chem. Soc. 134(19), 8022–8025 (2012). [CrossRef]  

6. Z. G. Xia, C. G. Ma, M. S. Molokeev, Q. L. Liu, K. Rickert, and K. R. Poeppelmeier, “Chemical unit cosubstitution and tuning of photoluminescence in the Ca2(Al1-xMgx)(Al1-xSi1+x)O7:Eu2+ phosphor,” J. Am. Chem. Soc. 137(39), 12494–12497 (2015). [CrossRef]  

7. Z. G. Xia, G. K. Liu, J. G. Wen, Z. G. Mei, M. Balasubramanian, M. S. Molokeev, L. C. Peng, L. Gu, D. J. Miller, Q. L. Liu, and K. R. Poeppelmeier, “Tuning of photoluminescence by cation nanosegregation in the (CaMg)x(NaSc)1-xSi2O6 solid solution,” J. Am. Chem. Soc. 138(4), 1158–1161 (2016). [CrossRef]  

8. G. G. Li, Y. Tian, Y. Zhao, and J. Lin, “Recent progress in luminescence tuning of Ce3+ and Eu2+−activated phosphors for pc-WLEDs,” Chem. Soc. Rev. 44(23), 8688–8713 (2015). [CrossRef]  

9. L. B. Su, J. Xu, H. J. Li, W. Q. Yang, Z. W. Zhao, J. L. Si, Y. J. Dong, and G. Q. Zhou, “Codoping Na+ to modulate the spectroscopy and photoluminescence properties of Yb3+ in CaF2 laser crystal,” Opt. Lett. 30(9), 1003–1005 (2005). [CrossRef]  

10. A. Pugžlys, G. Andriukaitis, A. Baltuška, L. Su, J. Xu, H. Li, R. Li, W. J. Lai, P. B. Phua, A. Marcinkevičius, M. E. Fermann, L. Giniūnas, R. Danielius, and S. Ališauskas, “Multi-mJ, 200-fs, cw-pumped cryogenically cooled, Yb,Na:CaF2 amplifier,” Opt. Lett. 34(13), 2075–2077 (2009). [CrossRef]  

11. 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, Q. Zhang, Y. X. Cao, and J. Xu, “Generation of 103fs mode-locked pulses by a gain linewidth-variable Nd,Y:CaF2 disordered crystal,” Opt. Lett. 39(7), 1737–1739 (2014). [CrossRef]  

12. H. Yu, D. P. Jiang, F. Tang, L. B. Su, S. Y. Luo, X. G. Yan, B. Xu, Z. P. Cai, J. Y. Wang, Q. W. Ju, and J. Xu, “Enhanced photoluminescence and initial red laser operation in Pr:CaF2 crystal via co-doping Gd3+ ions,” Mater. Lett. 206, 140–142 (2017). [CrossRef]  

13. P. W. Metz, D. T. Marzahl, G. Huber, and C. Kränkel, “Performance and wavelength tuning of green emitting terbium lasers,” Opt. Express 25(5), 5716–5724 (2017). [CrossRef]  

14. X. Liu, K. Yang, S. Zhao, T. Li, C. Luan, X. Guo, B. Zhao, L. Zheng, L. Su, J. Xu, and J. Bian, “Growth and lasing performance of a Tm,Y:CaF2 crystal,” Opt. Lett. 42(13), 2567–2570 (2017). [CrossRef]  

15. W. W. Ma, X. B. Qian, J. Y. Wang, J. J. Liu, X. W. Fan, J. Liu, L. B. Su, and J. Xu, “Highly efficient dual-wavelength mid-infrared CW laser in diode end-pumped Er:SrF2 single crystals,” Sci. Rep. 6(1), 36635 (2016). [CrossRef]  

16. C. Ziolek, H. Ernst, G. F. Will, H. Lubatschowski, H. Welling, and W. Ertmer, “High-repetitio-rate, high-average-power, diode pumped 2.94-µm Er:YAG laser,” Opt. Lett. 26(9), 599–601 (2001). [CrossRef]  

17. J. F. Zhu, L. Wei, W. L. Tian, J. X. Liu, Z. H. Wang, L. B. Su, J. Xu, and Z. Y. Wei, “Generation of sub-100fs pulses from mode-locked Nd,Y:SrF2 laser with enhancing SPM,” Laser Phys. Lett. 13(5), 055804 (2016). [CrossRef]  

18. 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]  

19. 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]  

20. Y. K. Voronko, A. A. Kaminskii, and V. V. Osiko, “Analysis of the optical spectra of CaF2:Nd3+ (type 1) crystals,” Sov. Phys. JETP 22(2), 295–300 (1966).

21. Y. K. Voronko, V. V. Osiko, and I. A. Shcherbakov, “Investigation of the interaction of Nd3+ ions in CaF2, SrF2 and BaF2 crystals (type 1),” Sov. Phys. JETP 28(5), 838–844 (1969).

22. 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]  

23. Y. V. Orlovskii, T. T. Basiev, V. V. Osiko, H. Gross, and J. Heber, “Fluorescence line narrowing (FLN) and site-selective fluorescence decay of Nd3+ centers in CaF2,” J. Lumin. 82(3), 251–258 (1999). [CrossRef]  

24. A. A. Kaminskii, V. V. Osico, 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]  

25. A. A. Kaminskii, “The Nature of “Aging” of CaF2-YF3-Nd3+ crystals (Type 1) under stimulated emission conditions,” Phys. Status Solidi B 20(1), K51–K54 (1967). [CrossRef]  

26. J. L. Doualan, L. B. Su, G. Brasse, A. Benayad, V. Ménard, Y. Y. Zhan, A. Braud, P. Camy, J. Xu, and R. Moncorgé, “Improvement of infrared laser properties of Nd:CaF2 crystals via codoping with Y3+ and Lu3+ buffer ions,” J. Opt. Soc. Am. B 30(11), 3018–3021 (2013). [CrossRef]  

27. C. R. A. Catlow and M. J. Norgett, “Shell model calculations of the energies of formation of point defects in alkaline earth fluorides,” J. Phys. C: Solid State Phys. 6(8), 1325–1339 (1973). [CrossRef]  

28. C. R. A. Catlow, A. V. Chadwick, G. N. Greaves, and L. M. Moroney, “Direction observations of the dopant environment in fluorite using EXAFS,” Nature 312(5995), 601–604 (1984). [CrossRef]  

29. E. L. Kitts, M. Ikeya, and J. H. Crawford, “Reorientation kinetics of dipolar complexes in Gadolinium-doped alkaline-earth fluorides,” Phys. Rev. B 8(12), 5840–5846 (1973). [CrossRef]  

30. 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]  

31. R. Iffländer, “Solid-state lasers for materials processing,” Springer-Verlag, pp. 311–318 (2001).

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

33. G. Kresse and J. Hafner, “Ab initio molecular-dynamics simulation of the liquid-metal-amorphous-semiconductor transition in germanium,” Phys. Rev. B 49(20), 14251–14269 (1994). [CrossRef]  

34. 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]  

35. G. Kresse and J. Furthmüller, “Efficient iterative schemes for ab initio total-energy calculations using a plane-wave basis set,” Phys. Rev. B 54(16), 11169–11186 (1996). [CrossRef]  

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

37. G. Kresse and D. Joubert, “From ultrasoft pseudopotentials to the projector augmented-wave method,” Phys. Rev. B 59(3), 1758–1775 (1999). [CrossRef]  

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

39. C. Persson, Y. J. Zhao, S. Lany, and A. Zunger, “n-type doping of CuInSe2 and CuGaSe2,” Phys. Rev. B 72(3), 035211 (2005). [CrossRef]  

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

41. 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]  

42. 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]  

43. 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]  

44. 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]  

45. C. R. A. Catlow, “The defect properties of anion-excess alkaline-earth fluorides:II intermediate and high dopant concentrations,” J. Phys. C: Solid State Phys. 9(10), 1859–1869 (1976). [CrossRef]  

46. F. Ma, D. Jiang, F. Tang, Q. Wu, S. Pang, Y. Liu, and L. Su, “Modulated photoluminescence parameters of neodymium in Sr0.95Y0.05F2.05 laser crystal,” Opt. Mater. Express 7(9), 3231–3237 (2017). [CrossRef]  

47. O. K. Alimov, T. T. Basiev, M. E. Doroshenko, P. P. Fedorov, V. A. Konyushkin, A. N. Nakladov, and V. V. Osiko, “Investigation of Nd3+ ions spectroscopic and laser properties in SrF2 fluoride,” Opt. Mater. 34(5), 799–802 (2012). [CrossRef]  

48. F. K. Ma, D. P. Jiang, L. B. Su, J. Y. Wang, W. Cai, J. Liu, J. G. Zheng, W. G. Zheng, J. Xu, and Y. Liu, “Spectral properties and highly efficient continuous-wave laser operation in Nd-doped Sr1-xYxF2+x crystals,” Opt. Lett. 41(3), 501–503 (2016). [CrossRef]  

49. M. Inokuti and F. Hirayama, “Influence of energy transfer by the exchange mechanism on donor luminescence,” J. Chem. Phys. 43(6), 1978–1989 (1965). [CrossRef]  

50. B. F. Aull and H. P. Jenssen, “Vibronic interactions in Nd:YAG resulting in nonreciprocity of absorption and stimulated emission cross sections,” IEEE J. Quantum Electron. 18(5), 925–930 (1982). [CrossRef]  

51. E. Friedman and W. Low, “Effect of thermal treatment of paramagnetic resonance spectra of rare earth impurities in calcium fluoride,” J. Chem. Phys. 33(4), 1275–1276 (1960). [CrossRef]  

Cited By

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

Alert me when this article is cited.


Figures (19)

Fig. 1.
Fig. 1. Thermodynamic stable Nd3+ centers in SrF2 crystals: (a) first group monomer centers, (b) second group clusters with cubic sublattice local environment and (c) third group clusters and the local is square antiprism. The bracket suffixed with different numbers means the same centers with different configurations.
Fig. 2.
Fig. 2. Thermodynamic stable Y3+ centers in SrF2 crystals: (a) first group monomer centers, (b) second group cubic sublattice clusters and (c) third group square antiprism clusters. The bracket suffixed with different numbers means the same centers with different configurations.
Fig. 3.
Fig. 3. Thermodynamic stable [Nd3+-Y3+] centers in SrF2 crystal: (a) second group cubic sublattice clusters, (b) third group square antiprism clusters. The bracket suffixed with different numbers means the same centers with different configurations.
Fig. 4.
Fig. 4. (a) Formation energy of Nd3+ and/or Y3+ clusters versus the number of Y3+ ions within a cluster. Hollow symbol with the same solid shape represents the same clusters with different configurations. (b) Number of dopant cations dependent formation energy of Nd3+ or Y3+ centers. (c) Local distortion of Nd3+ in [Nd3+−Y3+] polyhedron, the arrow with directions to Nd3+ and F means shortened and elongated bond length, respectively, when compared with the pure Nd3+ centers.
Fig. 5.
Fig. 5. Projected DOS on local coordination structure of the cluster. The p state of fluorine in [1Y3+−1Fi] and [1Nd3+−1Fi] centers with interstitial Fi at the nearest site (a) and (c), and the next nearest site (b) and (d), respectively. The insert represents p, d and f states of the corresponding rare-earth ions.
Fig. 6.
Fig. 6. Strategy of local structure manipulation. With no loss of crystal structure integrity, the cubic first coordination shell of active Nd3+ (blue core shell) would be manipulated to be distorted lower symmetric square antiprism sublattice when more than one Y3+ ion (green core shell) was introduced.
Fig. 7.
Fig. 7. Nominal molar concentration of Nd3+ dependent peak absorption cross section (a) and normalized PL intensity at 1057 nm (b) of y at.% Nd3+:SrF2 and y at.% Nd3+,5 at.% Y3+:SrF2. (c) Y3+ concentration dependent peak absorption cross section of 0.5 at.% Nd3+, x at.% Y3+:SrF2. (d) PL spectra of 0.5 at.% Nd3+, x at.% Y3+:SrF2, the intensity of sample A was multiplied by 1.5 for clear view. The insert is Y3+ concentration dependent normalized emission intensity at 1057 nm. Instrumental: SBW = 8 cm−1. The samples in same dimensions were measured under the same conditions, and the emission intensity could therefore be compared with each other.
Fig. 8.
Fig. 8. Low temperature TRES on clustering process. 3D image of the TRES of 4F3/24I11/2 transition of A (a), C (b), and F (c) samples. The insert represents the viewing along z-axis.
Fig. 9.
Fig. 9. Projected DOS of p state of interstitial Fi in the square antiprism lattice of (a) Y3+ clusters and (b) Nd3+ centers.
Fig. 10.
Fig. 10. Projected DOS of d state of Sr atom nearest to [1Nd3+−1Fi] center with Fi at the nearest site.
Fig. 11.
Fig. 11. XRD patterns of the crystal.
Fig. 12.
Fig. 12. Nominal Nd3+ concentration dependent product of peak absorption cross section (σabs) and FWHM (λabs) of 4I9/24F5/2 + 2H9/2 transition.
Fig. 13.
Fig. 13. Absorption cross section of the crystal, all peaks were attributed to the absorption of Nd3+ ions, the strongest band is around 800 nm corresponding to 4I9/24F5/2 + 2H9/2.
Fig. 14.
Fig. 14. Logarithmic decaying curves of the sample, for A, B, C and D crystals the lifetime consists of a long and short component, while for E and F the decaying could be well fitted by single exponential expression.
Fig. 15.
Fig. 15. Nominal Y3+ concentration dependent FWHM of 4F3/24I11/2 transition.
Fig. 16.
Fig. 16. Nominal Y3+ concentration dependent absolute quantum yields of 0.5 at.% Nd3+, x at.% Y3+:SrF2 crystal.
Fig. 17.
Fig. 17. Nominal Y3+ concentration dependent fluorescence and radiative lifetime of 0.5 at.% Nd3+, x at.% Y3+:SrF2.
Fig. 18.
Fig. 18. Nominal Y3+ concentration dependent peak emission cross section of 0.5 at.% Nd3+, x at.% Y3+:SrF2 crystal.
Fig. 19.
Fig. 19. 2D-TRES was ten sliced, integrated and normalized spectra of A (a), C (b), and F (c) crystal.

Tables (2)

Tables Icon

Table 1. Simulated formation energy of Nd3+ or Y3+ centers in SrF2 crystals. The bracket suffixed with different numbers means the same centers with different configurations, “ / ” denotes that the center was not stable.

Tables Icon

Table 2. Simulated formation energy of [Nd3+-Y3+] clusters in SrF2 crystal. The bracket suffixed with different numbers means the same centers with different configurations.

Equations (2)

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

Δ E = ( E t o t + E 0 ) m E 1 n E 2 w E 3 [ m + n + w ( m + n w ) 2 ] E c o r r
E c o r r = ( 1 + g ) q 2 α 2 ε L
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.