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

Computational model for operation of 2 µm co-doped Tm,Ho solid state lasers

Open Access Open Access

Abstract

A computational model for operation of co-doped Tm,Ho solid-state lasers is developed coupling (i) 8-level rate equations with (ii) TEM00 laser beam distribution, and (iii) complex heat dissipation model. Simulations done for Q-switched ≈0.1 J giant pulse generation by Tm,Ho:YLF laser show that ≈43 % of the 785 nm light diode side-pumped energy is directly transformed into the heat inside the crystal, whereas ≈45 % is the spontaneously emitted radiation from 3F4, 5I7, 3H4 and 3H5 levels. In water-cooled operation this radiation is absorbed inside the thermal boundary layer where the heat transfer is dominated by heat conduction. In high-power operation the resulting temperature increase is shown to lead to (i) significant decrease in giant pulse energy and (ii) thermal lensing.

©2007 Optical Society of America

1. Introduction

Co-doped Tm, Ho solid-state lasers present significant interest for a number of advanced applications. In recent years significant progress in understanding the basic phenomena underlying Tm,Ho laser operation as well as in developing high power lasers has been achieved [122]. However, many related physical effects which are non-linearly coupled with each other continue to remain unclear. For instance, notwithstanding the importance attributed to thermal effects [23], they are often treated in an excessively simplistic way, which does not allow a correct interpretation of experimental data as well as an adequate laser simulation and optimization. In this Communication we develop a coupled themo-optoical computational model in which specific non-steady-state thermal effects are rigorously coupled with population dynamics and spectroscopic processes involved in the energy transitions, and lasing in Tm,Ho solid state lasers. These effects are found to play a very significant role in high-power high- frequency Q-switched laser operation required for development of coherent-detection lidar systems.

2. Optics model

There are several models describing the electron population dynamics of co-doped Tm,Ho solid state lasers using the main levels involved in the pumping and laser generation shown in Fig. 1. These models have included various effects such as end and side pumping, ground state depletion, energy transfer between Tm3+ and Ho3+ ions, and also different types of up-conversion processes decreasing the upper laser level population. Several studies have used simplifications allowing model reduction to two-rate equations describing the electron density at the excited levels 3F4 and 5I7, and ground-state levels 3H6 and 5I8. In our study we endeavor to retain all the terms in the rate dynamics model, allowing us to reveal several specific effects. In particular, our analysis is based on computational non-steady state thermo-optical model using 8-level rate equations and related data describing the population dynamics of Tm,Ho lasers from Walsh et al. [17]:

dn1dt=Rp(t)+n2τ2+p28n2n8p71n7n1p41n4n1+p22n22,
+p27n2n7p51n5n1p61n6n1+p38n3n8
dn2dt=n2τ2+n3τ3p28n2n8+p71n7n1+2p41n4n12p22n22,
p27n2n7+p51n5n1
dn3dt=n3τ3+n4τ4+p61n6n1p38n3n8,
dn4dt=Rp(t,z,r)n4τ4p41n4n1p22n22,
dn5dt=n5τ5+p27n7n2p51n5n1,
dn6dt=n6τ6+n5τ5p61n6n1p38n8n3.

For the upper laser level (5I7):

dn7dt=n7τ7+n6τ6p28n2n8p71n7n1p27n2n7+p51n5n1cσseη(f7n7f8n8)ϕ(t,r).

For the lower laser level (5I8):

dn8dt=n7τ7p28n2n8+p71n7n1+p61n6n1p38n3n8cσseη(f7n7f8n8)ϕ(t,r),

where n i(t,r) are the level concentrations, p ij are the probabilities of the optical transitions, τ i are the level lifetimes, R p(t) is the pumping source, ϕ(t,r) is the local laser photon density, σse is the stimulated emission cross-section, f i(t,r) are the Boltzmann level population factors and η is the refractive index of the crystal.

 figure: Fig. 1.

Fig. 1. Energy transfer processes in co-doped Tm,Ho materials and energy differences used in Eq. (16).

Download Full Size | PDF

All optical transition probabilities and level lifetimes, including characteristic radiative times, have been considered in detail in Ref. [17]. Although that study neglects some of the possible radiation decays from the upper manifolds which could easily be taken into account in present study, we have not extended the original model of Ref. [17] in view of the good agreement with the known experimental data on Q-switched pulse operation [6].

The local laser photon density ϕ(t,r) is represented by the product of (i) the total number of photons inside the oscillator cavity, Φ0(t), depending on t and (ii) the normalized space distribution function, ϕ 0(r). The resulting equation for Φ0(t) is given by a differential equation including integration of the stimulated and spontaneous radiation over the crystal volume [17, 23, 24]:

dΦ0(t)dt=Φ0(t)cσseηVcr(f7n7f8n8)ϕ0(r,z)dVΦ0(t)τc+ετ7Vcrn7dV,

where V cr is the crystal volume, and ε≈10-7-10-8 is a factor taking into account the proportion of photons spontaneously emitted within the solid angle of the mirrors, and τ c is the cavity lifetime given by:

τc1=c2Lopt[lnR1ln(1Tout)+β],

where L opt=L cav+(η-1)L cr is the characteristic optical length, L cav is the cavity length and L cr is the crystal length; R l is the back mirror reflectance, T out is the output mirror transmittance and β is the parameter used in our simulations for the optical loss associated with the active Q-switching: β=0 for the open resonator and β≫-ln R 1(1-T out) for the closed resonator. For the acousto-optic Q-switch, if the fraction of the main beam diffracted out of the resonator is 0.9, β=-ln(1-0.9)=2.3. We neglect here additional reflectance and scattering loss on crystal and Q-switch. However, these factors can also be included into the round trip optical loss in Eq. (10).

For the case of 100–-500 ns pulse generation considered here the cavity length L cavL cr and the spatial photon distribution inside the operating crystal can be described by TEM00 fundamental mode as:

ϕ0(r)=2πw02Lcavexp(2r2w02),

where w 0 is the beam waist radius of TEM00 mode defined by the resonator parameters.

The solution of the rates equations together with the main oscillator Eq. (9) gives the radial distribution of the output power density (W/m2) at the output mirror as:

I0(t,r)=Φ0chνlas2Loptln11Tout×2πw0*2exp(2r2w0*2),

where w*0 is the modified beam radius outside the resonator (for instance, for the case of a TEM00 Gaussian beam inside the confocal spherical resonator one has w0=Lcavλl2π and w*0=√2w 0 at the output mirror).

In this paper we consider a particular case of Tm (6%), Ho(0.4%):YLF operation side pumped by 785 nm LD radiation. For 6 % Tm doped YLF crystal one finds for the absorption coefficient α=σ a N Tm≈2.8 cm -1 [17]. Thus, a 2 mm diameter YLF crystal is able to absorb a (1-exp(-2αd))≈0.67 of the incident beam flux in the case of the double-pass pumping scheme providing high uniformity of the absorbed flux over the crystal volume. In the simulation we assume that the laser rod axis is directed along the c-axis of the YLF crystal and we neglect anisotropy in absorption of the polarized beams of the LD bars. We assume that this anisotropy is not significant for the side-pumped configuration in which three LD bars are arranged around the crystal in threefold symmetry and the internal surface of the tube used for water cooling has a high diffusive reflection [25]. In fact, high incident fluxes are able to deplete the 3H6-level in Tm3+ and to reduce significantly absorption [23]. However, in this study the concentration of the 3H6-level does not fall below 0.9 of the Tm-concentration, the related variations of α do not exceed 5 %, and in the simulations we use:

Rp(t)ηpηaQpπd2LcrhνpΔtp×{1,tΔtp0,t>Δtp.

where η a=(1-ρ)[1-exp(-2 αd)] is the absorption efficiency of pumping, ρ is the reflection factor of the pumping radiation into laser material, Q p is the pumping pulse energy, Δt p is the pumping pulse duration and η p is quantum efficiency.

In Fig. 2 we show a simulated giant pulse (G-pulse) generated by a Tm,Ho:YLF laser producing ≈0.1 J pulses of ≈150 ns duration. In particular, we simulate an active Q-switched laser side-pumped by 0.5 ms LD pulses of 785 nm wavelength for a crystal 2 cm long and 2 mm in diameter placed inside a 1 m long cavity (T out=0.05 and R l=0.98) with a 0.85 mm radius waist in the TEM00 laser beam distribution. The Q-switch is open after a 0.5 ms pumping period with a delay of 0.7 ms to ensure that the G-pulse generation starts after achieving the maximal possible gain. This delay is associated with the delay of excitation transfer from 3H4 to 3F4, and finally towards the lasing 5I7 level [22].

 figure: Fig. 2.

Fig. 2. Simulation of G-pulse generation: pulse power versus time.

Download Full Size | PDF

3. Thermal model

The heat absorbed inside the crystal leads to a temperature increase over the crystal volume. For high power operation this temperature shift is able to change the local values of the Boltzmann population factors of the upper and lower lasing levels:

fi(t,r)=giexp[EikBT(t,r)]jgjexp[EjkBT(t,r)],

where k B is the Boltzmann constant, g i is the degeneracy of the i-level, and T (t,r) is the local temperature.

Generally, the operating crystal is heated via lattice vibrations due to non-radiative decay of electrons from all levels involved in the excitations. The local heat source is defined by:

qcr(t,r)=i=27ΔEiniτinr,

where ΔE i is the energy difference between the i-manifold and the next lower manifold into which the electron makes the transition (Fig. 1) and τ inr are the non-radiative times inversely proportional to the non-radiative transition probabilities.

In order to avoid difficulties in defining the probabilities of non-radiative transitions, an estimate of the heat source can be made via the difference between the pumped energy and the energy of stimulated and spontaneous radiation leaving the crystal [12]. This approach is mainly used for the CW mode or as an averaged estimate for high-repetition pulsed mode. However, we use this approach for normal or Q-switched mode operation by introducing a modification which takes into account the rate, i=27ΔEi*dnidt, at which the pumped energy is stored inside Tm3+ and Ho3+ ions as:

qcr(t,r)=Rp(t)hνpcσseη1hνl(f7n7f8n8)ϕ(t,r)i=27ΔEiniτiri=27ΔEi*dnidt,

where in addition to ΔE i we introduce the energy difference between the i-manifold and the ground state ΔE*i (Fig. 1), and then τ ir are the corresponding radiative times [17].

The calculation of Eq. (16) for the Tm,Ho:YLF laser reveals several effects significant for energy extraction by lasing pulse. First, Fig. 3(a) shows the energy balance integrated over the crystal volume versus time. It reveals a very significant extension of the heat release period as compared with the pumping period. Fig. 3(a) shows that the heat is released inside the crystal over a period of ≈10 ms, whereas the pumping period is 0.5 ms during which only ≈30 % of heat is released. A two-time lower resulting temperature increase is achieved in the crystal prior to G-pulse generation (1.2 ms) as follows from an estimate neglecting the thermal conductivity effect:

 figure: Fig. 3.

Fig. 3. Energy balance versus time during laser operation: (a) energy pumping and release and (b) optical loss by spontaneous radiation from different levels.

Download Full Size | PDF

ΔTcr(t,r)1ρc0tqcr(t,r)dt.

Second, Fig. 3(a) also shows that in the final energy balance ≈0.12 J corresponds to the G-pulse energy, ≈0.75 J corresponds to the heat released inside the crystal and ≈0.84 J corresponds to the energy lost by spontaneous emission. Thus, about 43 % of the pumped energy is directly converted into heat. We should note that the estimates of heat release based on 2-level rate equations treat this value as the difference between the pumped energy and the optical energy of the laser pulse and the spontaneous emission from two levels, 3F4 and 5I7 [12]. The energy spontaneously emitted by other levels, i.e. 3H5, 3H4, 5I5 and 5I6, are implicitly included into the heat released inside the crystal [12]. The 8-level model used here shows that the contribution of the 5I5 and 5I6 levels into the spontaneous emission loss is negligibly small, whereas the contribution of 3H4 and 3H5 appears to be quite significant, ≈0.1 and ≈0.4 J, respectively (see Fig. 3(b)). Adding these values to the heat release of ≈0.75 J gives ≈70 %, similar to the result from the 2-level model [12].

Thus, only ≈43 % of the pumped energy is released directly as heat inside the crystal, whereas ≈45 % is spontaneously emitted radiation from the crystal at wavelengths: λ2=1.93 µm, λ 3=4.32 µm, λ 4=2.46 µm and λ 7=2.07 µm. These wavelengths are within the transparency range of the crystal and are therefore able to leave the crystal. Fig. 3(b) shows the values of the total optical loss, i=27ΔEiniτir, and also the losses emitted from all levels, ΔE i n i/τ ir integrated over the crystal volume. These radiation fluxes leaving the crystal are absorbed by the water flow typically used for crystal cooling. The water absorption coefficients for these wavelengths are [26]: α 2=124 cm-1, α 3=300 cm-1, α 4=63.5 cm-1 and α 7=31 cm-1. That is, the spontaneously emitted fluxes leaving the crystal are absorbed within lengths of ≈α -1 i, i.e. within 80, 33, 157 and 320 µm from the surface, respectively. The absorption of these fluxes in the vicinity of the crystal surface can significantly inhibit heat dissipation from the crystal. The heat transfer to the water flow depends on the Reynolds number, Re, defining the level of the flow turbulency dependent on the water flow rate through the channel inside which the operating crystal is set up. Numerical estimates show that for the typical coaxial crystal in a tube water channel geometry and typical flow rates, the value of the heat transfer coefficient is h=103-105 W/m2 K [27]. The main thermal resistance to the heat flow from the crystal surface is due to the thermal boundary layer, δ T, within which the heat conductance dominates over the convective transport. The estimate of δ T follows from the equivalency of -k crT cr/∂r|sur=-k wT w/∂r|sur=h(T cr|sur-T w∞), where k cr≈6 and k w≈0.6 W/m K are the thermal conductivity of crystal and water, respectively. That is, using ∂T w/∂r|sur≈-(T cr|sur-T w∞)/δ T one finally obtains for h=103-105 W/m2K:

δTkwh6600μm.

Thus, the spontaneous IR fluxes are absorbed by water within a distance where the heat transfer is dominated by the thermal conductivity. Hence, the absorption of these fluxes is able to significantly inhibit the heat dissipation from the crystal. In order to consider the thermal effect we simulate the complex heat transfer non-steady state, two-dimensional problem by coupling the above optical model with the heat generation and heat transport through the operating crystal, and the water boundary layer inside which the absorption of spontaneously emitted IR radiation takes place. The radially symmetric temperature distribution inside the cylindrical crystal, T cr(t, r), and the thermal boundary layer in water, T w(t, r), are defined by:

ρiCiTit=(kiTi)+qi(t,r),

for crystal (i=cr) and water (i=w) with the boundary condition T w=T w∞ at r=R 0+δ T, where δ T=R 0[exp(k w/R 0 h)-1] takes into account the radial curvature.

Heat source density inside the crystal is defined by Eq. (16) whereas the heat source density due to the absorption of spontaneously emitted IR fluxes in water is defined by:

qw(t,r)=R0riJ0i(t)αiexp[αi(rR0)],

where J 0i(t) are the IR flux densities isotropically leaving the crystal given by:

J0i(t)=1ScrVΔEini(t,r)τirdV.

The effect of IR radiation absorption is negligibly small for h>105 W/m2K and δ T<6 µm, when δ Tα -1 i. However, for h≈104 W/m2K (δ T≈60 µm) this effect is very significant, and can lead to the onset of an inverted temperature distribution inside the crystal when the temperature inside the boundary layer is higher than that inside the crystal. This effect is shown to take place in a coupled thermo-optical simulation performed using a conservative 50x100 conservative finite-difference approximation. The main results of this simulation given in Fig. 4 at three times show that a significant temperature increase has occured (≈2 K) by the start of G-pulse generation (1.2 ms), which leads to a pulse energy decrease due to the decrease in f 7 n 7-f 8 n 8 over the crystal volume. This figure also shows that an inverted temperature distribution inside the crystal is present at period of time of ≈10 ms.

 figure: Fig. 4.

Fig. 4. Temperature distribution inside the operating Tm,Ho:YLF crystal and thermal boundary layer for single G-pulse generation for h=104 W/m2K (δ T≈60 µm, water temperature T w=290 K).

Download Full Size | PDF

This effect, associated with the strong absorption of emitted IR radiation in water, produces a different result in high repetition mode. In particular, the simulation of 20 and 50 Hz G-pulse repetition mode given in Fig. 5 shows that in contrast with first few pulses after pulsed operation stabilization the temperature inside the crystal becomes higher than that at the crystal surface due to the onset of a quasi-steady state gradient. Finally, Fig. 5(a) shows that this thermal effect leads to ≈10–25 % reduction of G-pulse energy combined with a strong thermal lensing effect known to be detrimental for laser beam quality.

4. Summary

A complex thermo-optical model for Tm,Ho solid state lasers has been developed based on an 8-level rate dynamics model for the excitation transfer to Ho3+ ions from LD pumped Tm3+ ions integrated together with the equation for the total number of stimulated photons inside the cavity. This model is also coupled with a two-dimensional time dependent heat transfer model including absorption, heat release and heat transfer inside the operating crystal as well as the absorption and the thermal effect of infrared radiation fluxes spontaneously emitted by the operating crystal. In the case of water cooled laser operation the thermal effect is shown to be split into two simultaneously occurring processes: (i) direct heat release inside the crystal and (ii) infrared spontaneously emitted radiation fully absorbed in water over a distance of several hundreds of microns, which corresponds to a typical value of boundary layer thickness. In particular, the simulations show that only ≈43 % of the pumped energy is transformed into heat directly inside the crystal, whereas ≈45 % is IR radiation spontaneously emitted by 3H4, 3H5, 3F4 and 5I7 levels and absorbed in the vicinity of the crystal surface. The absorption taking place within the boundary layer provides an additional strong thermal effect, inhibiting the dissipation of the heat from the crystal and significantly increasing crystal temperature. The resulting temperature increase is shown to reduce significantly G-pulse energy.

 figure: Fig. 5.

Fig. 5. 20 and 50 Hz G-pulse laser operation: (a) G-pulse power modification with time and (b) temperature increase in the operating crystal versus time for h=104 W/m2 K for crystal axis and surface.

Download Full Size | PDF

Acknowledgments

We would like to acknowledge the financial support from the National Institute of Information and Communications Technology (Japan). We would also like to thank Dr. J. Hester from the Australian Nuclear Science and Technology Organization for careful reading of this paper and valuable comments.

References and links

1. J.K. Tyminski, D.M. Franich, and M. Kokta “Gain dynamics of Tm,Ho:YAG pumped in near infrared,” J. Appl. Phys . 65, 3181–3188 (1989). [CrossRef]  

2. V.A. French, R.R. Petrin, R.C. Powell, and M. Kokta, “Energy-transfer processes in Y3Al5O12:Tm,Ho,” Phys. Rev . B 46, 8018–8026 (1992).

3. R.R. Petrin, M.G. Jani, R.C. Powell, and M. Kokta, “Spectral dynamics of laser-pumped Y3Al5O12:Tm,Ho lasers,” Opt. Mater . 1,111–124 (1992). [CrossRef]  

4. M.G. Jani, R J. Reeves, R.C. Powell, G.J. Quarles, and L. Esterovitz, “Alexandrite-laser excitation of a Tm:Ho:Y3Al5O12 laser.” J. Opt. Soc. Am . B 8, 741–746 (1991).

5. M. G. Jani, F. L. Naranjo, N. P. Barnes, K.E. Murray, and G.E. Lockard, “Diode-pumped long-pulse-length Ho:Tm:YLiF4 laser at 10 Hz,” Opt. Lett . 20, 872–874 (1995). [CrossRef]   [PubMed]  

6. J. Yu, U.N. Singh, N.P. Barnes, and M. Petros “125-mJ diode-pumped injection-seeded Ho:Tm:YLF laser,” Opt. Lett . 23, 780–782 (1998). [CrossRef]  

7. A.N. Alpat’ev, V.A. Smirnov, and I.A. Shcherbakov, “Relaxation oscillations of the radiation from a 2-µm holmium laser with a Cr,Tm,Ho:YSGG crystal,” Quantum Electron . 28, 143–146 (1998). [CrossRef]  

8. I.F. Elder and M.J.P. Payne, “Lasing in diode-pumped Tm:YAP, Tm,Ho:YAP and Tm,Ho,YLF,” Opt. Commun . 145, 329–339 (1995). [CrossRef]  

9. N. P. Barnes, E. D. Filer, C. A. Morrison, and C. J. Lee, “Ho:Tm Lasers I: Theoretical,” IEEE J. Quantum Electron . 32, 92–103 (1996). [CrossRef]  

10. C. J. Lee, G. Han, and N.P. Barnes, “Ho:Tm Lasers II: Experiments,” IEEE J. Quantum Electron . 32, 104–111 (1996). [CrossRef]  

11. G. Rustad and K. Stenersen, “Modeling of laser-pumped Tm and Ho lasers accounting for upconversion and ground-state depletion,” IEEE J. Quantum Electron . 32, 1645–1656 (1996). [CrossRef]  

12. D. Bruneau, S. Delmonte, and J. Pelon, “Modeling of Tm,Ho:YAG and Tm,Ho:YLF 2-µm lasers and calculation of extractable energies,” Appl. Opt . 37, 8406–8419 (1998). [CrossRef]  

13. G. L. Bourdet and G. Lescroart, “Theoretical modeling and design of a Tm,Ho:YLiF4 microchip laser,” Appl. Opt . 38, 3275–3281 (1999). [CrossRef]  

14. S. D. Jackson and T.A. King, “CW operation of a 1.064-µm pumped Tm-Ho-Doped silica fiber laser,” IEEE J. of Quantum Electron . 34, 1578–1587 (1998). [CrossRef]  

15. V. Sudesh and K. Asai, “Spectroscopic and diode-pumped-laser properties of Tm,Ho:YLF; Tm,Ho:LuLF; and Tm,Ho:LuAG crystals: a comparative study,” J. Opt. Soc. Am . B 20, 1829–1837 (2003).

16. A. Sato, K. Asai, and K. Mizutani, “Lasing characteristics and optimizations of diode-side-pumped Tm,Ho:GdVO4 laser,” Opt. Lett . 29, 836–838 (2004). [CrossRef]   [PubMed]  

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

18. G. Galzerano, E. Sani, A. Toncelli, G. Della Valle, S. Taccheo, M. Tonelli, and P. Laporta, “Widely tunable continuous-wave diode-pumped 2-µm Tm-Ho:KYF4 laser,” Opt. Lett . 29, 715–717 (2004). [CrossRef]   [PubMed]  

19. J. Izawa, H. Nakajima, H. Hara, and Y. Arimoto, “Comparison of lasing performance of Tm,Ho:YLF lasers by use of single and double cavities,” Appl. Opt . 39, 2418–2421 (2000). [CrossRef]  

20. J. Yu, B. C. Trieu, E. A. Modlin, U.N. Singh, M. J. Kavaya, S. Chen, Y. Bai, P. J. Petzar, and M. Petros, “1 J/pulse Q-switched 2 µm solid-state laser,” Opt. Lett . 31, 462–464 (2006). [CrossRef]   [PubMed]  

21. X. Zhang, Y. Ju, and Y. Wang, “Theoretical and experimental investigation of actively Q-switched Tm,Ho:YLF lasers,” Opt. Express 14, 7745–7750 (2006). [CrossRef]   [PubMed]  

22. O.A. Louchev, Y. Urata, and S. Wada, “Numerical simulation and optimization of Q-switched 2 µm Tm,Ho:YLF laser,” Opt. Express 15, 3940–3947 (2007). [CrossRef]   [PubMed]  

23. P. Černý and D. Burns, “Modeling and experimental investigation of a diode-pumped Tm:YAlO3 laser with a- and b-cut crystal orientations,” IEEE J. of selected topics in quantum electron . 11, 674–681 (2005). [CrossRef]  

24. V.P. Risk, “Modeling of longitudinally pumped solid-state lasers exhibiting reabsorption losses,” J. Opt. Soc. Am . B 5, 1412–1423 (1988).

25. D. Golla, M. Bode, S. Knoke, W. Schöne, and A. Tünnermann, “62-W cw TEM00 Nd:YAG laser side-pumped by fiber-coupled diode lasers,” Opt. Lett . 21, 210–212 (1996). [CrossRef]   [PubMed]  

26. D. M. Wieliczka, S. Weng, and M. R. Querry, “Wedge shaped cell for highly absorbent liquids: infrared optical constants of water,” Appl. Opt . 28, 1714–1719 (1989). [CrossRef]   [PubMed]  

27. W. Koechner, Solid-State Laser Engineering, 6th Edition (New-York, Springer, 2006).

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 (5)

Fig. 1.
Fig. 1. Energy transfer processes in co-doped Tm,Ho materials and energy differences used in Eq. (16).
Fig. 2.
Fig. 2. Simulation of G-pulse generation: pulse power versus time.
Fig. 3.
Fig. 3. Energy balance versus time during laser operation: (a) energy pumping and release and (b) optical loss by spontaneous radiation from different levels.
Fig. 4.
Fig. 4. Temperature distribution inside the operating Tm,Ho:YLF crystal and thermal boundary layer for single G-pulse generation for h=104 W/m2K (δ T ≈60 µm, water temperature T w=290 K).
Fig. 5.
Fig. 5. 20 and 50 Hz G-pulse laser operation: (a) G-pulse power modification with time and (b) temperature increase in the operating crystal versus time for h=104 W/m2 K for crystal axis and surface.

Equations (23)

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

dn 1 dt = R p ( t ) + n 2 τ 2 + p 28 n 2 n 8 p 71 n 7 n 1 p 41 n 4 n 1 + p 22 n 2 2 ,
+ p 27 n 2 n 7 p 51 n 5 n 1 p 61 n 6 n 1 + p 38 n 3 n 8
dn 2 dt = n 2 τ 2 + n 3 τ 3 p 28 n 2 n 8 + p 71 n 7 n 1 + 2 p 41 n 4 n 1 2 p 22 n 2 2 ,
p 27 n 2 n 7 + p 51 n 5 n 1
dn 3 dt = n 3 τ 3 + n 4 τ 4 + p 61 n 6 n 1 p 38 n 3 n 8 ,
dn 4 dt = R p ( t , z , r ) n 4 τ 4 p 41 n 4 n 1 p 22 n 2 2 ,
dn 5 dt = n 5 τ 5 + p 27 n 7 n 2 p 51 n 5 n 1 ,
dn 6 dt = n 6 τ 6 + n 5 τ 5 p 61 n 6 n 1 p 38 n 8 n 3 .
dn 7 dt = n 7 τ 7 + n 6 τ 6 p 28 n 2 n 8 p 71 n 7 n 1 p 27 n 2 n 7 + p 51 n 5 n 1 c σ se η ( f 7 n 7 f 8 n 8 ) ϕ ( t , r ) .
dn 8 dt = n 7 τ 7 p 28 n 2 n 8 + p 71 n 7 n 1 + p 61 n 6 n 1 p 38 n 3 n 8 c σ se η ( f 7 n 7 f 8 n 8 ) ϕ ( t , r ) ,
d Φ 0 ( t ) dt = Φ 0 ( t ) c σ se η V cr ( f 7 n 7 f 8 n 8 ) ϕ 0 ( r , z ) dV Φ 0 ( t ) τ c + ε τ 7 V cr n 7 dV ,
τ c 1 = c 2 L opt [ ln R 1 ln ( 1 T out ) + β ] ,
ϕ 0 ( r ) = 2 π w 0 2 L cav exp ( 2 r 2 w 0 2 ) ,
I 0 ( t , r ) = Φ 0 ch ν las 2 L opt ln 1 1 T out × 2 π w 0 * 2 exp ( 2 r 2 w 0 * 2 ) ,
R p ( t ) η p η a Q p π d 2 L cr h ν p Δ t p × { 1 , t Δ t p 0 , t > Δ t p .
f i ( t , r ) = g i exp [ E i k B T ( t , r ) ] j g j exp [ E j k B T ( t , r ) ] ,
q cr ( t , r ) = i = 2 7 Δ E i n i τ inr ,
q cr ( t , r ) = R p ( t ) h ν p c σ se η 1 h ν l ( f 7 n 7 f 8 n 8 ) ϕ ( t , r ) i = 2 7 Δ E i n i τ ir i = 2 7 Δ E i * dn i dt ,
Δ T cr ( t , r ) 1 ρ c 0 t q cr ( t , r ) dt .
δ T k w h 6 600 μ m .
ρ i C i T i t = ( k i T i ) + q i ( t , r ) ,
q w ( t , r ) = R 0 r i J 0 i ( t ) α i exp [ α i ( r R 0 ) ] ,
J 0 i ( t ) = 1 S cr V Δ E i n i ( t , r ) τ ir dV .
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.