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

Broad-spectrum and long-lifetime emissions of Nd3+ ions in lead fluorosilicate glass

Open Access Open Access

Abstract

A novel Nd3+-doped lead fluorosilicate glass (NPS glass) is prepared by a two-step melting process. Based on the absorption spectrum a Judd-Ofelt theory analysis is made. The emission line width of NPS glass is 44.2nm. The fluorescence decay lifetime of the 4F3/2 level is 586±20µsec, and the stimulated emission cross-section is 0.87×10-20cm2 at 1056nm. A laser oscillation is occurred at 1062nm when pumped by 808nm Diode Laser. The slope efficiency is 23.7% with a 415mJ threshold. It is supposed that NPS glass is a good candidate for using in ultra-short pulse generation and amplification by the broad emission bandwidth and long fluorescence lifetime.

©2009 Optical Society of America

1. Introduction

High-intensity ultra-short-pulse laser, which can generate some extreme physical conditions in aspects of temperature, pressure, electric/magnetic field, energy density and so on [1], is one of the most powerful research tools in many major scientific subjects, including fast ignition (FI) for laser fusion, generation of high brightness X-ray, laboratory astrophysics, advanced accelerator techniques, exploration of the effects between strong light and materials, etc.

In general, the ultra-short-pulse laser is developed mainly in two directions, one is broad-spectrum ultra-short-pulse, and another is large energy and high peak power. Both are interrelated. Some negative factors such as gain-narrowing of spectrum, nonlinear optical phenomena and optical damage are restricting the development of large-energy, high-power and narrow-pulse laser. For the most popular used Ti: sapphire, although it has deeply promoted the commercialization of femto-second laser and atto-second pulse output, it cannot produce ultra-short-pulse laser with large energy due to small crystal size, narrow bandwidth and short fluorescence lifetime. Novel laser medium with large size and high performance such as broad spectrum is required [2, 3].

Glass is an attractive host because that depending on the composition, the glasses can be highly transparent from UV to IR and the solubility of the rare earth ions is higher compared to that in crystals. Furthermore, glass is easy to be shaped and processed, especially to be formed in large size. Presently, high-power Nd3+-doped glass laser drivers are being used to generate large-energy short-pulse laser in some world-famous facilities such as OMEGA in United States and “SHEN-GUANG” in China.

In the traditional silicate and phosphate laser glasses, emission bandwidths of Nd3+ are about 25nm [4], and a laser pulse with a duration of 50fs can be produced under the diffraction limit condition. However, more short duration can hardly be obtained because of the negative effects of spectral gain narrowing and dispersion. For enhancing the spectral width of Nd3+-doped glass, some approaches are performed. One direct method is to form a glass by mixing both silicate and phosphate components. But since the discrepancy of emission peaks in Nd3+-doped silicate and phosphate laser glasses is about 10nm, the broadening of transmission bandwidth is limited [3].

Another method is to choose wide-bandwidth laser glass as gain medium. Due to the low phonon energy, good fluorescence spectra and long lifetimes, the heavy metal oxide glasses (such as tantalum silicate glasses [5], Bi2O3-PbO-Ga2O3 [6], phospho-tellurite [7]), heavy metal fluoride glasses [8], oxyfluoride glass [9] and fluorophosphate glass [10] have been paid close extensive attention.

In this paper, a novel Nd3+-doped lead fluorosilicate glass (NPS glass) prepared by a two-step melting process was reported. The fluorescence bandwidth obtained in this glass is more than 40nm and the decay lifetime of Nd3+ for the 4F3/2 level is up to 586±20µsec.

2. Experimental procedure

The NPS glass composition studied in this work was (in mol %): 67SiO2-10PbO-2PbF2-12K2O-8Na2O-0.76Nd2O3-0.24As2O3. The high-purity (ࣙ99.99%) raw materials were added in the form of SiO2, Pb3O4, PbF2, KNO3, NaNO3, Nd2O3 and As2O3 powders. The contents of the most common transition metal ions are less than 5ppm and that of the physically or chemically absorbed water is less than 0.1wt%. A mixture of required amounts (typically 1.9kg) of raw materials was carefully grinded, thoroughly mixed and pre-melted in a quartz crucible at 1250 °C. At this temperature, the first melting can be carried out quickly with low-level fluoride evaporation. After melted into vitreous state, the mixture was manually transferred into a Φ95mm×100 mm high-purity platinum metal crucible with lid, then refined at 1420 °C to remove bubbles and homogenized at 1380 °C by platinum stirrer for 3h separately. The molten glass was removed out at 1360 °C and cast into a copper mold to form plates, which had been heated to 350 °C beforehand. Finally the glass was annealed in a custom-designed annealing oven to reduce residual stress. It was kept for 24h at 470 °C just above the glass transition temperature, T g, firstly and then cooled to room temperature at a rate of about 10 °C/h.

The glass was cut and polished for optical and spectroscopic measurements. These samples have fine performances such as stria-free, the parallelism with 1 minutes of arc, the surface flatness with <λ/4 at 633nm in the 90% area of central aperture, surface quality with 40-20 scratch-dig.

All optical and spectroscopic measurements were carried out at room temperature. The refractive indices were measured using a WYV V-Prism Refractometer from Shanghai Precision & Scientific Instrument Co., Ltd with accuracy of ±5×10-5. Refractive index dispersion curve was fitted with the Sellmeier formula using the values of refractive indices at six wavelengths including 435.8, 486.1, 546.1, 589.3, 656.3 and 706.5nm. The absorption spectrum in the range of 400–1000nm was recorded with a Lamda 950 UV/VIS/NIR spectrophotometer from Perkin-Elmer. The resolution was set to 1nm and the precision for the various measured quantities is as follows: wavelength (±0.08nm) and optical transmission (±0.1%). The luminescent spectrum was determined using a spectrophotometer (JobinYvon Triax320, France), with an 808nm laser diode (LD) (Coherent, U.S.A) as the excitation source, and the power is 400mw. The fluorescence decay rates were measured by changing the excitation source to pulsed output, with the pulse-width and period are 20μs and 5000μs, respectively. The signals were detected by an InGaAs photodetector (J22TE2-66C-R05M-1.7, Judson) in the range of 800–1650nm wavelength. The temporal decay of the fluorescence signal was recorded by a storage digital oscilloscope (Tektronix TDS3012B, U.S.A). Errors in the fluorescence and lifetime measurements are estimated to be ±0. 1nm, ±20µs, respectively.

To perform the laser oscillation, a resonator was applied with the parameter R1=100%, R2=90%, respectively. An 808nm semiconductor laser array with the pulse duration of 200 µs at a repetition rate of 1 Hz was used as the pumping. The laser output energy is measured by Rj-7620 with RjP-735 Pyroelectric Probe from Laser Probe Inc. with a calibration uncertainty of ±5%. The laser spectrum is carried out by HR2000+ High-resolution Spectrometer from Ocean Optics Inc. with an optical resolution of 0.05nm.

3. Results and discussion

3.1 Application of Judd-Ofelt theory

The fitted refractive index dispersion curve and the absorption spectrum of NPS glass are shown in Fig. 1. The refractive index decreases with the increase of light’s wavelength, which follows as typical dispersion curves of heavy metal fluoride glasses. The absorption bands of Nd3+ ions in NPS glass are parallel to those of commercial Nd3+-doped aluminum-phosphate laser glass LG-770 (Schott Glass Technologies, Inc.), except for lower absorption cross sections. Eight absorption bands centering at 431, 473, 531, 574, 684, 740, 807 and 881 nm are ascribed to 4I9/22P1/2+2D5/2, 4I9/22 K 15/2+2G9/2+4G11/2+(2D,2F)3/2, 4I9/24G9/2+2K13/2 +4G7/2, 4I9/22G7/2+4G5/2+2H11/2, 4I9/24F9/2, 4I9/24S3/2+4F7/2, 4I9/24F5/2+2H9/2 and 4I9/24F3/2 transitions, respectively.

The Judd-Ofelt theory was used to calculate the radiative rates of Nd3+ ions in the NPS glass. According to the Judd-Ofelt theory [11, 12], the experimental oscillator strength, fmeas, can be calculated by the following equation,

fmeas=mc2πe2N0κ(λ)dλλ2

where m and e are the mass and charge of the electron respectively, c is the vacuum velocity of light, N 0 is the number of rare-earth ions per unit volume in the glass, and the k(λ) is absorption coefficient at a given wavelength λ, which is given by

κ(λ)=2.303log(I0I)l

where l is the thickness of glass, I 0 andIare the intensities of incident and emergent lights, respectively.

 figure: Fig. 1.

Fig. 1. Room temperature optical absorption spectra of Nd3+-doped NPS glass (size: 10mm×10mm×3mm) and commercial Nd3+-doped aluminum-phosphate laser glass LG-770. All transitions start from 4I9/2 level to the indicated levels. The fitted refractive index dispersion curve of the NPS glass (red in color) is also shown.

Download Full Size | PDF

The theoretical oscillator strengths, fcal, are given by electric dipole oscillator strength, fcaled, magnetic-dipole oscillator strengths, fcalmd, and electric-quadrupole, fcaleq, where the values of fcalmd (~10-9) and fcaleq (~10-11) are much smaller in comparison to that of fcaled (~10-6), therefore can be neglected generally [11, 12]. Thus, the fcal is expressed by the equation,

fcal=fcaled=8π2mc3hλ(2J+1)×(n2+2)29n×t=2,4,6Ωt4fN(SL)JU(t)4fN(SL)J2

where J is the total angular momentum of the initial level, which for Nd3+ is 9/2, and (n 2+2)2/9 is the local field correction using the tight binding approximation. The reduced matrix elements of the unit tensor operators, <‖U(t)‖>, are calculated in the intermediate-coupling approximation. They are found to be almost invariant to the environment and are given by Carnall et al. [13].

The oscillator strengths both experimentally and theoretically obtained are presented in Table 1. The three Judd-Ofelt intensity parameters, Ωt(t=2, 4, 6) are therefore obtained by the least-square fitting of the theoretical oscillator strength values to the measured ones. The quality of the fitting can be expressed by the root-mean-square, δ rms, which is calculated by the following formula,

δrms=[(fcalfmeas)2NbandsNp]

where Nbands regards the number of transitions bands analyzed, Np is 3 in this case, which is the number of Qt parameters.

The spontaneous emission probability, A (J′→J″), for excited levels of rare earth ions from an initial state J′ to a final ground state J″ can be identified as

A(JJ)=64π4e23h(2J+1)λp3·n(n2+2)29t=2,4,6Ωt4F32U(t)4IJ2

By using the Ωt, values and the matrix elements <‖U (t)‖> for Nd3+ obtained from the literature [14] the probabilities for emission from 4 F 3/2 to the 4 IJ″(J″=9/2, 11/2, 13/2, 15/2) manifolds were calculated. These probabilities are listed in Table 2. The Ωt values are used to calculate the radiative transition probabilities, branching ratios and radiative lifetimes.

Tables Icon

Table 1. Measured and calculated oscillator strengths for NPS glass

Tables Icon

Table 2 . Calculated radiative transition rates AJ, luminescence branching ratios β and corresponding radiative lifetime τrad for Nd3+ in NPS glass

The Nd3+ fluorescence bands are asymmetric, therefore an effective line-width, Δλeff, is defined by

Δλeff=I(λ)dλImax

where I max is the maximum intensity at each fluorescence emission peak.

The radiative lifetime can be shown as

τrad=1JArad(J,J)

The branching ratio can be written as

β(4F324IJ)=A(4F324IJ)JA(4F324IJ)

Finally, the emission cross section, σemi, can be determined by the following relation

σemi=λp48πcn2ΔλeffA[(4F324IJ)]

The calculated Judd-Ofelt intensity parameters and many important spectroscopic characteristics in this work and earlier reports [5, 6, 7, 8, 9, and 15] are summarized in Table 3.

Tables Icon

Table 3 . Optical Properties of NPS glass

It is currently accepted that the parameter Ω 2 exhibits the dependence on the covalency between Nd3+ ions and ligand anions, and the Ω 2 also reflects the asymmetry of the local environment at the Nd3+ ions site [10, 16]. In this way, the weaker the value of Ω 2 is, the more centre-symmetrical of the ion site and ionic of its chemical bond with the ligands are. This indicates that Nd3+ ions in NPS glass present low covalence and asymmetry in bonding.

Emission intensity could be expressed in terms of the ratio Ω46 because Ω2 is not included to calculate branching ratios for the laser 4 F 3/24 I 11/2 transition of Nd3+ ions. The value of Ω46 for the NPS glass is 0.86. This value is near the range of 0.9–1.1, which implies the good fluorescent performance [17].

As shown in Table 3, the value of the stimulated emission cross section for the 4F3/24I11/2 transition is 0.87×10-20cm2 at the peak emission wavelength. The relatively small cross section for the NPS glass arises in part from smaller intensity parameters Ω6 and from the fluorescence bandwidth which is much broader than those of the similar systems illustrated in Table 3. In addition, containing of larger sized potassium in NPS glass composition is also blamed to induce small emission cross section [17].

3.2 Fluorescence spectra and Radiative properties

 figure: Fig. 2.

Fig. 2. (a) Room temperature emission spectrum of NPS glass under diode laser excitation at 808nm. (b) Nd3+ energy level diagram for Nd3+ doped NPS glass obtained from room temperature absorption and emission spectra. (c) Luminescence decay curve of 4F3/2 level of Nd3+ ions in NPS glass. (The sample sizes: 5mm×5mm×1mm)

Download Full Size | PDF

Fluorescence spectrum of NPS glass under 808nm LD measured at room temperature is shown in Fig.2 (a). Three major emission bands were observed at 880, 1055 and 1325nm, corresponding to the 4F3/24I9/2, 4F3/24I11/2 and 4F3/24I13/2 transitions, respectively. Fig.2 (b) shows Nd3+ energy level diagram in Nd3+-doped NPS glass obtained from its absorption and emission spectra. Fig. 2(c) also shows the recorded luminescence decay curve of Nd3+ for the most characteristic peak of 4F3/24I11/2 transition. Analysis of the decay curve results in a long lifetime of 586±20µs. The value is much longer than those of other glass systems reported before (as summarized in Table 3). According to Equ.5 and 7, the radiative lifetime of rare-earth ions is strongly depending on the refractive index, J-O parameters Ω 4 and Ω 6. Compared to other host glasses labeled in Table 3 and Fig. 3, the NPS glass has relatively small refractive index, J-O parameters Ω 4 and Ω 6, therefore presents longer radiative lifetime.

Figure 3 shows the emission spectrum of NPS glass at room temperature. It can be seen that NPS glass has much broader effective line-width than Nd3+: Bi2O3-PbO-Ga2O3 [6], Phosphotellurite [7], and Phosphate (LG-770) [15] glass systems, except for tantalum silicate glasses (K-824) [5].

 figure: Fig. 3.

Fig. 3. Emission spectra for 4F3/24I11/2 level of Nd3+ ion in laser glasses at room temperature (The sample sizes: 5mm×5mm×1mm).

Download Full Size | PDF

The broad effective line-width of 1056nm emission in NPS glass derives from two possible factors. Firstly, in NPS glass, Pb2+ ion in the [PbO4] tetragonal pyramid has a pair inert electron in outer shell, thus remains large polarizability [18]. Secondly, the amount of ionic bonds in NPS glasses is relatively high due to the substitution of F- for O2- in part. The asymmetry of [NdO6] octahedron is therefore large. These two factors induce the Stark splitting of Nd3+ ion’s J″ manifolds and result in the inhomogeneous broadening of emission spectra. Relatively large value of Ω 2 for NPS glass also proved the high-level asymmetry of [NdO6] octahedron in NPS glass.

The peak wavelength of fluorescence spectra has a tight relation with the basic glass former network [17]. The NPS glass is adjudged to lead silicate system, and the fluorescence peak wavelength is therefore similar to those in the silicate glasses. The introduction of F can make a little blue-shift of fluorescence peak wavelength [9]. So, the emission linewidth of NPS glass is much broader than those of Nd3+-doped laser glasses reported in previous works and the fluorescence lifetime is also much longer. The peak emission cross section is small compared with that of most materials, but the emission lifetime and effective bandwidth is unique among Nd3+-doped materials, thereby permitting the efficient extraction of energy from an Nd3+-doped NPS glass amplifier [2, 4].

3.3 Laser output property

In the laser oscillation measurement, laser output was observed with increasing the input pump energy. The observed laser emission spectrum is displayed in Fig. 4(a), together with the measured threshold and slope efficiency curve. It can be seen that the laser peak was centered at the wavelength of 1062±0.05nm and the full width at half maximum of the laser spectrum above threshold was 0.6±0.05nm. While in Fig. 4 (b), along with the enhancement of pumping power, a sharp bend in that curve was observed indicating a threshold value of approximately 415±21mJ. The maximum laser output was 32mJ and the slope efficiency was determined to be 23.7±1.2%.

 figure: Fig. 4.

Fig. 4. (a) Laser oscillation at 1062 nm in the Nd3+-doped NPS glass. (b) Measured threshold and slope efficiency. (The sample sizes: 20mm×20mm×10mm)

Download Full Size | PDF

4. Conclusion

A novel Nd3+-doped lead fluorosilicate glass 67SiO2-10PbO-2PbF2-12K2O-8Na2O-0.76Nd2O3-0.24As2O3 (in mol %) was prepared by a two-step melting process. The optical properties were investigated by means of optical absorption, luminescence and lifetime measurements. The Judd-Ofelt parameters were calculated and the radiative rates, lifetimes and branching ratios were obtained. Laser output was achieved with a wavelength of 1062nm. The laser threshold was 415mJ and the slope efficiency was 23.7%. The broad emission bandwidth and long fluorescence lifetime make NPS-glass a good candidate for uses in ultra-short pulse generation and amplification.

Acknowledgment

This research was financially supported by the National Natural Science Foundation of China (NSFC, No.60808023, No. 10876009) and one Hundred Talents Programs of the Chinese Academy of Sciences.

References and links

1. Z. Xu and R. Li, “Recent progress in the development of high intensity ultrashort pulse lasers at SIOM,” Chin. Opt. Lett. 5, S1–S4 (2007), http://www.opticsinfobase.org/col/abstract.cfm?URI=col-5-101-S1.

2. G. Chériaux and J. Chambaret, “Ultra-short high-intensity laser pulse generation and amplification.” Meas. Sci. Technol.12, 1769–1776 (2001), http://dx.doi.org/10.1088/0957-0233/12/11/303.

3. G. R. Hays, E. W. Gaul, M. D. Martinez, and T. Ditmire, “Broad-spectrum neodymium-doped laser glasses for high-energy chirped-pulse amplification.” Appl. Opt. 46, 4813–4819 (2007), http://www.opticsinfobase.org/ao/abstract.cfm?URI=ao-46-21-4813. [CrossRef]   [PubMed]  

4. Z. Jiang, J. Yang, and S. Dai, “Optical spectroscopy and gain properties of Nd3+-doped oxide glasses,” J. Opt. Soc. Am. B 21, 739–743 (2004), http://www.opticsinfobase.org/abstract.cfm?URI=josab-21-4-739. [CrossRef]  

5. S. E. Stokowski, R. A. Saroyan, and M. J. Weber, “Nd-Doped Laser Glass, Spectroscopy, and Physical Properties.” M-95, Rev. 2, 1 and 2 (Lawrence Livermore National Laboratory, University of California, Livermore, CA, 1981).

6. L. Kassab, S. Tatumi, C. Mendes, L. Courrol, and N. Wetter, “Optical properties of Nd doped Bi2O3-PbO-Ga2O3 glasses,” Opt. Express 6, 104–108 (2000), http://www.opticsinfobase.org/oe/abstract.cfm?URI=oe-6-4-104. [CrossRef]   [PubMed]  

7. M. J. Weber, J. D. Myers, and D. H. Blackburn, “Optical properties of Nd3+ in tellurite and phosphotellutite glasses.” J. Appl. Phys. 52, 2944–2949 (1981), http://link.aip.org/link/?JAPIAU/52/2944/1. [CrossRef]  

8. J. Azkargorta, I. Iparraguirre, R. Balda, and J. Fernández, “On the origin of bichromatic laser emission in Nd3+-doped fluoride glasses,” Opt. Express 16, 11894–11906 (2008), http://www.opticsinfobase.org/oe/abstract.cfm?URI=oe-16-16-11894. [CrossRef]   [PubMed]  

9. M. Abril, J. Méndez-Ramos, I. R. Martín, U. R. Rodríguez-Mendoza, V. Lavín, A. Delgado-Torres, and V. D. Rodríguez. “Optical properties of Nd3+ ions in oxyfluoride glasses and glass ceramics comparing different preparation methods,” J. Appl. Phys. 95, 5271–5279 (2004), http://link.aip.org/link/?JAPIAU/95/5271/1. [CrossRef]  

10. Ju H. Choi, Alfred Margaryan, Ashot Margaryan, Frank G. Shi, and Wytze Van Der Veer, “Fluorescence and Nonradiative Properties of Nd3+ in Novel Heavy Metal Contained Fluorophosphate Glass,” Advances in OptoElectronics 2007 , 1–8 (2007), http://dx.doi.org/10.1155/2007/39892. [CrossRef]  

11. B. R. Judd, “Optical absorption intensities of rare earth ions,” Phys. Rev. 127, 750–761 (1962). [CrossRef]  

12. G. S. Ofelt, “Intensities of crystal spectra of rare-earth ions,” J. Chem. Phys. 37, 511–520 (1962). [CrossRef]  

13. W. T. Carnall, P. R. Fields, and B. G. Wybourne, “Spectral Intensities of the Trivalent Lanthanides and Actinides in Solution. I. Pr3 +, Nd3 +, Er3 +, Tm3 +, and Yb3 +.” J. Chem. Phys. 42, 3797–3806 (1965), http://link.aip.org/link/?JCPSA6/42/3797/1

14. A. A. Kaminskii, G. Boulon, M. Buoncristiani, B. Di Bartolo, A. Kornienko, and V. Mironov, “Spectroscopy of a new laser garnet Lu3Sc2Ga3O12:Nd3+. Intensity luminescence characteristics, stimulated emission, and full set of squared reduced-matrix elements |<‖U(t)‖>|2 for Nd3+ ions.” Phys. Status Solidi A 141, 471–494 (1994). [CrossRef]  

15. J. H. Campbell, P. Ehrmann, T. Suratwala, W. Steele, C. Thorsness, and M. McLean, “Properties of and Manufacturing Methods for NIF Laser Glasses.” Inertial Confinement Fusion Quarterly Report 9, 118–119 (1999).

16. E. O. Serqueira, N.O. Dantas, A. F. G. Monte, and M. J. V. Bell. “J,dd Ofelt calculation of quantum efficiencies and branching ratios of Nd3+ doped glasses.” J. Non-Crystal. Solids 352, 3628–3632 (2006), http://dx.doi.org/10.1016/j.jnoncrysol.2006.03.093. [CrossRef]  

17. R. R. Jacobs and M. J. Weber, “Dependence of the 4F3/2→4I11/2 induced-emission cross section for Nd3+ on glass composition.” IEEE J. Quantum Electron. QE-12, 102–111(1976). [CrossRef]  

18. M. B. Saisudha and J. Ramakrishna, “Effect of host glass on the optical absorption properties of Nd3+, Sm3+, and Dy3+ in lead borate glasses.” Phys. Rev. B 53, 6186–6196 (1996), http://prola.aps.org/pdf/PRB/v53/i10/p6186_1. [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 (4)

Fig. 1.
Fig. 1. Room temperature optical absorption spectra of Nd3+-doped NPS glass (size: 10mm×10mm×3mm) and commercial Nd3+-doped aluminum-phosphate laser glass LG-770. All transitions start from 4I9/2 level to the indicated levels. The fitted refractive index dispersion curve of the NPS glass (red in color) is also shown.
Fig. 2.
Fig. 2. (a) Room temperature emission spectrum of NPS glass under diode laser excitation at 808nm. (b) Nd3+ energy level diagram for Nd3+ doped NPS glass obtained from room temperature absorption and emission spectra. (c) Luminescence decay curve of 4F3/2 level of Nd3+ ions in NPS glass. (The sample sizes: 5mm×5mm×1mm)
Fig. 3.
Fig. 3. Emission spectra for 4F3/24I11/2 level of Nd3+ ion in laser glasses at room temperature (The sample sizes: 5mm×5mm×1mm).
Fig. 4.
Fig. 4. (a) Laser oscillation at 1062 nm in the Nd3+-doped NPS glass. (b) Measured threshold and slope efficiency. (The sample sizes: 20mm×20mm×10mm)

Tables (3)

Tables Icon

Table 1. Measured and calculated oscillator strengths for NPS glass

Tables Icon

Table 2 Calculated radiative transition rates AJ, luminescence branching ratios β and corresponding radiative lifetime τrad for Nd3+ in NPS glass

Tables Icon

Table 3 Optical Properties of NPS glass

Equations (9)

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

fmeas=mc2πe2N0 κ(λ)dλλ2
κ(λ)=2.303log(I0I)l
fcal=fcaled=8π2mc3hλ(2J+1)×(n2+2)29n×t=2,4,6Ωt4fN(SL)JU(t)4fN(SL)J2
δrms=[(fcalfmeas)2NbandsNp]
A(JJ)=64π4e23h(2J+1)λp3·n(n2+2)29t=2,4,6Ωt4F32U(t)4IJ2
Δλeff=I(λ)dλImax
τrad=1JArad(J,J)
β(4F324IJ)=A(4F324IJ)JA(4F324IJ)
σemi=λp48πcn2ΔλeffA[(4F324IJ)]
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.