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

Flowing-gas diode pumped alkali lasers: theoretical analysis of transonic vs supersonic and subsonic devices

Open Access Open Access

Abstract

We examine transonic diode pumped alkali laser (DPAL) devices as a simpler alternative to supersonic devices, suggested by B.D. Barmashenko and S. Rosenwaks [Appl. Phys. Lett. 102, 141108 (2013)], where complex hardware, including supersonic nozzle, diffuser and high power mechanical pump, is required for continuous closed cycle operation. Three-dimensional computational fluid dynamics modeling of transonic (Mach number M ~0.9) Cs and K DPALs, taking into account the kinetic processes in the lasing medium is reported. The performance of these lasers is compared with that of supersonic (M ~2.5) and subsonic (M ~0.2) DPALs. For Cs DPAL the maximum achievable power of transonic device is lower than that of supersonic, with the same resonator and Cs density at the laser section inlet, by only ~3% implying that supersonic operation mode has only small advantage over transonic. On the other hand, for subsonic laser the maximum power is by 7% lower than in transonic, showing larger advantage of transonic over subsonic operation mode. The power achieved in supersonic and transonic K DPALs is higher than in subsonic by ~80% and ~20%, respectively, showing a considerable advantage of supersonic device over transonic and of transonic over subsonic.

© 2016 Optical Society of America

1. Introduction

Diode pumped alkali lasers (DPALs) are populated via the D2(n2S1/2 → n2P3/2) transition of the alkali atoms, followed by rapid relaxation (by buffer gas, helium and\or small hydrocarbon molecules) of the upper to the lower fine-structure level, n2P3/2 to n2P1/2 (designated as levels 3 and 2, respectively; the ground state n 2S1/2 is designated as 1), resulting in D1(n2P1/2→ n2S1/2) lasing [1,2]. In spite of the high efficiency of DPALs, high pumping power results in heat release, mainly due to the relaxation, and leads to temperature rise and decrease of the power and efficiency of the laser. Obviously, flowing the gas in the laser cell enhances the heat transfer out of the cell, impedes temperature rise and increases the power and efficiency of the laser.

Demonstration of subsonic DPALs [3,4] and modeling of both supersonic and subsonic devices [5–15] taking into account fluid dynamics and kinetic processes in the lasing medium, show the positive influence of the gas flow on the laser performance, where the highest lasing power and optical-to-optical efficiency were predicted for supersonic DPALs. In the present paper we examine transonic DPALs as a simpler alternative to supersonic devices, where complex hardware, including supersonic nozzle, diffuser and high power mechanical pump (needed for recovery of the gas total pressure which strongly drops in the diffuser), is required for continuous closed cycle operation [6] (the term “transonic” is chosen merely to imply high subsonic velocity and does not impose a new flow regime). The laser power Plase of transonic lasers is compared with that of supersonic and subsonic lasers for optimal pressures, temperatures and buffer gas compositions in the laser section found in [5,6]. Pressure variations have the strongest influence on the laser power. On the one hand the rate of relaxation of n2P3/2 increases with increasing pressure, resulting in increase of Plase. On the other hand the rate of absorption on the D2 pump transition decreases with increasing pressure for narrowband pumping (with bandwidth smaller than the absorption linewidth) resulting in decrease of Plase. The trade-off between these two processes determines the optimal pressure (several atmospheres). For subsonic and supersonic lasers we used the same values of the Mach number M as in [5,6], (0.2 and ~2.5, respectively), whereas for transonic laser M = 0.9 was chosen, to be in the high subsonic flow regime while still sufficiently far from sonic, M = 1, conditions where transition to local supersonic may occur.

As discussed in [1,2], flowing-gas DPALs are of interest due to their scalability potential to very high power. In the present paper we compare different devices, with rather large cross sections of the flow, 5 x 5 cm2, and of the pump beam, 2 x 5 cm2, in order to learn if megawatt (MW) laser power can be achieved in them. The main conclusions of the comparison are valid also for lower pumping power (where the dimensions of the device are correspondingly smaller).

2. Description of the model

Schematic of a transonic device with transverse pumping dealt with in this paper is depicted in Fig. 1(a). Gas mixture of X/M (X = Cs, K and M = CH4 or He, respectively) flows in a converging subsonic nozzle with A/Ai = H/Hi ~1.6, where A and Ai are the cross section areas of the nozzle inlet and exit, respectively, and H and Hi are the corresponding nozzle heights. A laser section of constant height Hi, width W and length L (in the flow direction x) is located downstream of the nozzle exit. Both the pump and laser beams fill the entire laser section volume. Due to different values of spin–orbit splitting ΔE, the relaxation of the n2P3/2 levels of K and Cs atoms is carried out by different buffer gases. For the relatively small ΔE = 57.7 cm−1 of K, efficient spin–orbit mixing can be provided by pure helium as a buffer gas at practical pressures ~1atm. However, for Cs atoms, where ΔE = 554.1 cm−1 is rather large (compared to K), hydrocarbons (methane and ethane) buffer gases are used for efficient relaxation. A drawback of these buffer gases is that they may react with Cs atoms, contaminate the gain medium and decrease the beam quality due their relatively high refraction index [1].

 figure: Fig. 1

Fig. 1 Schematics of the transonic (a), supersonic (b) and subsonic (c) devices with transverse pumping by rectangular pump beam of 5 x 2 cm2 aperture. The flow is perpendicular to the laser and pump beams propagation axes. Note that different buffer gases are used for Cs and K.

Download Full Size | PDF

For comparing the transonic DPAL performance with that of subsonic and supersonic, the same alkali number density, nX,i = 8.4 × 1013 cm−3, at the laser section inlet is assumed. Based on this density and Mach number at the laser section inlet, Mi, the inlet temperature is calculated as follows. The values of Ti, pi and nX,i at the inlet to the laser section are related to their respective stagnation values (with the subscript “stag”) by the isentropic flow equations [16]:

TiTstag=(pipstag)γ1γ=(nX,innX,stag)γ1=(1+γ12Mi2)1,
where γ is the ratio of heat capacities at the operating temperatures of ~400 −550 K, γHe = 1.67, γCH4 = 1.26-1.20 and Ti, pi are the temperature and pressure at the laser section inlet, respectively. We assume that the temperature of the alkali source is equal to the stagnation temperature, Tstag, of the flow and thus find Ti which satisfies the condition nsat(Tstag) > nX,stag where nsat(Tstag) is the saturated vapor number density at the vapor source. This condition is necessary to prevent alkali vapor nucleation. The saturated alkali number density is:
nsat=psatkBTstag10abTstagkBTstag.
where psat is the saturated vapor pressure given in the usual exponential form with constants a and b according to the alkali in question and kB is Boltzmann constant [1]. As follows from Eq. (1) the alkali number density at the stagnation conditions is:
nX,stag=nX,i(1+γ12Mi2)1γ1
Tstag is found from the solution of the transcendental equation nsat = nX,stag, Eqs. (2) and (3), and the corresponding Ti is calculated from Eq. (1). M at the nozzle inlet can also be calculated from the nozzle area ratio A/Ai and M at the nozzle exit. At the nozzle inletA/A*=(A/Ai)(Ai/A*)M=0.9, whereA*is the critical cross section of the nozzle corresponding to M = 1. Using the isentropic flow relations [16] one finds for both DPALs studied, K and Cs, that the nozzle inlet M is ~0.4. For given Mi, Ti, and nX,i optimal values of pi corresponding to the maximum output power were used in the calculations (this point is dealt with in the Introduction).

To avoid condensation of X, the nozzle walls are heated to a higher temperature than that of the saturated X vapor. The laser resonator consists of a reflector and an output coupler with reflectivity r1 (close to 100%) and r2, located outside the laser cell. The schematic of the supersonic and subsonic devices also studied in this paper is presented in Figs. 1(b) and 1(c), respectively.

The model considers transverse flow in the x direction perpendicular to the laser optical axis z and transverse pumping configuration where the pump beams with spectral bandwidth ∆νp propagate perpendicularly to the optical axis z in the y direction (Fig. 1).

The flow and beam parameters in the laser section are determined from a three-dimensional computational fluid dynamics (3D CFD) model based on an ANSYS commercial code that calculates the gas dynamics and chemistry in the active medium. The flow-field is computed by a set of 3D conservation Eqs. (-) mass and momentum conservation for the pressure and velocity fields, and energy conservation for the gas temperature field. The molecular transport and chemical kinetics are defined by the species conservation equations, where the species represent the different energy states of the alkali atoms. The transport equations and the kinetics processes in K and Cs are presented in more detail in [6,15].

The computational mesh for the resonator section consists of nx x ny x nz hexahedral cells (nx = 30, ny = 80, nz = 80) and is shown in Fig. 2. The flow boundary conditions were the mass flow rate at the “flow in” and the pressure outlet at the “flow out” boundaries. The boundary conditions allow for the required flow velocity, gas temperature, mixture composition and gas pressure in the medium. For the walls we assumed constant temperature.

 figure: Fig. 2

Fig. 2 The mesh for the rectangular transverse-pumping beam, consists of nx x ny x nz hexahedral cells (nx = 30, ny = 80, nz = 80).

Download Full Size | PDF

The model of Fabry-Perot resonator, often employed in flowing-gas lasers with wide mirror aperture [17] was applied for calculating the laser intensity in the resonator. The laser beam is divided into nx x ny very narrow beams propagating independently in the optical axis direction z. The intensity of each part of the rectangular beam depends on both the x and y coordinates. The reason is that the local gain coefficients for the pump and laser beams are not uniform in their cross sections, resulting in non-uniform distributions of the pump and laser intensities. Similarly to the laser beam, the pump beams are also divided into nx x nz very narrow beams propagating independently in the y direction. The pump and laser intensities are described by the Beer-Lambert law [6,15]. In the case of broadband pumping the spectral distribution of the pump intensity Ip±(x,y,z,v) over the frequency v should be taken into account; the superscripts + and – indicate the pump beams propagating in the opposite directions + z and –z. The rates of changes of Ip±(x,y,z,ν) with z are given by

Ip±(x,y,z,ν)y=±g31(x,y,z,ν)Ip±(x,y,z,ν),
where g31(x,y,z,ν) is the local gain coefficient (usually negative) of the pump D2 transition at a location (x, y, z) given by:
g31(x,y,z,ν)=σ31(x,y,z,ν)[n3(x,y,z)2n1(x,y,z)],
where
 σ31(x,y,z,ν)=σ31(x,y,z,νp)f31(ν)
is the stimulated emission cross section for pumping the D2 transition at frequency v ,
 σ31(x,y,z,ν)=σ31,atΔν31,nΔν31(x,y,z),
ni is the number density of the alkali specie at energy level i = 1, 2, 3 obtained from the kinetic equations as described in [6,15],
f31(ν)=(Δν31/2)2/[(ν νp)2+(Δν31/2)2]     
and νp is the pump frequency.

In Eq. (7) σ31(x,y,zp) and σ31,at are the stimulated emission and atomic cross section, respectively, in the line center and Δν31,n and Δν31(x, y, z) are the natural and the pressure broadened widths of the D2 transition, respectively, where the spatial variations of σ31(x,y,zp) are caused mainly by the variations of the gas temperature. In Eq. (8), f31(ν) is the pressure broadened Lorentzian spectral distribution.

The boundary condition for Ip+at z = 0 and Ipat z = W are

Ip±(ν)= PpWLtfp(ν),
where Pp is the pump power, t is the windows transmission and
fp(ν)=(4ln2πΔνp2)1/2exp[4ln2(ννpΔνp)2]
is the pump normalized Gaussian spectral distribution of full width at half maximum (FWHM) Δνp [15].

The laser intensities in both directions Il±(x,y,z), where z = 0 is assumed to be located at the inner surface of the window closer to the output coupling mirror with reflectivity r2, are assumed to be monochromatic and their spatial variations are described by equations similar to (4):

Il±(x,y,z)z=±g21(x,y,z)Il±(x,y,z),
where g21(x,y,z) is the local gain coefficient of the laser D1 transition given by:
g21(x,y,z)=σ21(x,y,z,νl )[n2(x,y,z)n1(x,y,z)],
where
σ21(x,y,z,νl)=σ21,atΔν21,nΔν21(x,y,z)
is the stimulated emission cross section for the lasing D1 transition at frequency νl, σ21,at the atomic cross section in the line center andΔν21,n and Δν21(x,y,z) are the natural and the pressure broadened widths of the D1 transition, respectively.

The boundary conditions for the two way laser intensities Il+ and Il at z = 0, were calculated by ray tracing of the given output laser flux,Ilase(x,y): Il(x,y,0)=Ilase(x,y)/[t(1r2)] and Il+(x,y,0)=Il(x,y,0)t2r2.The solution for the output laser flux was found by iterating Ilase until Il(x,y,z=W)/Il+(x,y,z=W)=t2r1. The output power Plase is given by

Plase=0<x<L,0<y<HIlase(x,y)dxdy.

3. Results and discussion

Figure 3 shows the dependence of Plase on the pump power, Pp, for subsonic (M ~0.2) transonic (M ~0.9) and supersonic (M ~2.7) Cs DPALs with CH4 as a buffer gas, and with optimal parameters that are summarized in Table 1. We found out that variations of M by ~ ± 10% for each of the flow regimes at a given temperature in the laser section, result in a very weak changes of Plase ~ ± 1%. Thus, calculations made for the chosen values of M show the main trends for the laser power dependence on the flow regime which are by and large independent of the exact values of M and depend mainly on pi and Ti.

 figure: Fig. 3

Fig. 3 Dependence of Plase on Pp for subsonic (M ~0.2) transonic (M ~0.9) and supersonic (M ~2.7) Cs DPALs with CH4 as a buffer gas, and with optimal parameters summarized in Table 1.

Download Full Size | PDF

Tables Icon

Table 1. Optimal parameters of different Cs DPALs with CH4 as a buffer gas, transversely pumped with 5 x 2 cm2 rectangular pump beam, with reflector r1 = 0.95, output coupler r2 = 0.35 and windows transmission t = 0.98. For transonic flow pstag = 10 atm, Tstag = 425K and for supersonic pstag = 70 atm, Tstag = 493 K.

Calculations were first performed for narrowband pumping of Cs with pump bandwidth ∆νp = 10 GHz (examples 1-3 in Table 1). The maximum achievable power of ~1.8 MW for the supersonic device obtained at Pp ~3.5 MW is higher than that of the transonic and subsonic devices by only ~3% and ~10%, respectively, and the value of the maximum optical-to-optical efficiency of ~80% is about the same for the three cases (cases 1-3 in Table 1). Thus, in the case of Cs DPALs the supersonic operation mode has only small advantage over transonic. At the same time for the subsonic laser the maximum power is by 7% lower than for the transonic, showing a larger advantage of the transonic over the subsonic operation mode. Extremely large ~60% losses of alkali atoms due to ionization at Pp = 3.5 MW result in substantial decrease of the power and the efficiency of the laser. The reason for these losses is a very high intra-resonator radiation intensity of Ip ~2 × 109 W/m2 near the pump beam entrances. The active medium temperature rise in both supersonic and transonic cases is negligible (~10 K) due to their high flow velocities of ~1160 m/s and ~450 m/s, respectively, and the main reason for the small difference between the curves is the lower inlet temperature of the former.

The maximum values of the pump power exceeding hundreds of kW and having narrowband spectrum are not easily achievable for the present state-of-the-art. That is why we also performed computations for more realistic case of broadband pumping (example 4 in Table 1). The results are shown in Fig. 3 where the dependence Plase(Pp) is presented for the transonic Cs DPAL with ∆νp = 200 GHz (case 4 in Table 1). For the transonic laser with ∆νp = 10 GHz, Plase is larger than that calculated for ∆νp = 200 GHz over the whole range of Pp, its maximum value ~1.75 MW achievable at Pp ~3.5 MW being larger than that calculated for ∆νp = 200 GHz by only ~10%. The reason for this relatively small difference is the high gas mixture pressure (6 atm) that results in increase of the pumping beam absorption due to the increase of the D2 transition spectral broadening.

Figure 4 shows the dependence Plase(Pp) for subsonic (M ~0.2) transonic (M ~0.9) and supersonic (M ~2.4) K DPALs with He as a buffer gas and with optimal parameters that were found in [5,6], and are summarized in Table 2. In this case the maximum achievable power of ~2.8 MW in the supersonic K DPAL is by more than 45% larger than ~1.9 MW obtained for the transonic laser with the same resonator and K vapor density at the inlet. These powers are higher than the power achievable in the subsonic laser, by ~80% and ~20%, respectively, showing considerable advantages of the supersonic device over the transonic and of the transonic device over the subsonic. The fact that the power increase for the K supersonic laser is much larger than for the Cs device is explained by a much smaller fine-structure splitting of the 2P state (~58 cm−1 for K and ~554 cm−1 for Cs). Indeed, for rapid relaxation between the fine structure levels 3 and 2 the ratio between the populations of these levels is given by Boltzmann distribution:

n3n2=2exp(ΔEkBT).
High T results in increased n3 and decreased n2 and thus in decreased pump absorption and lasing rate, i.e., low optical-to-optical efficiency. The influence of the temperature on the population ratio is stronger for smallerΔE, resulting in a much stronger effect of the temperature decrease in the supersonic K DPAL.

 figure: Fig. 4

Fig. 4 Dependence of Plase on Pp for subsonic (M ~0.2) transonic (M ~0.9) and supersonic (M ~2.4) К DPALs with He as a buffer gas, and with optimal parameters summarized in Table 2.

Download Full Size | PDF

Tables Icon

Table 2. Optimal parameters of different K DPALs with He as a buffer gas, transversely pumped with 5 x 2 cm2 rectangular pump beam, with reflector r1 = 0.95, output coupler r2 = 0.35 and windows transmission t = 0.98; the pump bandwidth Δvp = 10 GHz. For transonic flow pstag = 11 atm, Tstag = 485K and for supersonic pstag = 90 atm, Tstag = 586 K.

4. Summary

Comparative 3D CFD modeling of supersonic, transonic and subsonic DPALs pumped by large cross section (5 x 2 cm2) beams, taking into account fluid dynamics and kinetic processes in the lasing medium, is reported. For Cs DPALs the supersonic operation mode has no substantial advantage over the transonic one, the main processes limiting the power being saturation of the D2 transition and large ~60% losses of alkali atoms due to ionization, whereas the influence of gas heating is negligible. At the same time for K DPALs the maximum value of the power in the supersonic laser is higher than that calculated for the transonic laser by more than 45%, showing a considerable advantage of the supersonic device over the transonic; the power of the transonic laser is by 20% higher than that of the subsonic laser showing an advantage of transonic device over subsonic. For broadband (~200 GHz) pumping of transonic Cs DPALs, the maximum achievable Plase is smaller than that for the narrowband (10 GHz) transonic case by only ~10%. The reason for this relatively small difference is the high optimal pressure of the gas mixture that increases the pumping beam absorption by increasing the D2 transition spectral broadening.

Acknowledgments

Effort sponsored by High Energy Laser Joint Technology Office (HEL-JTO) and the European Office of Aerospace Research and Development (EOARD) under grant FA9550-15-1-0489 and by Israel Science Foundation (ISF) under grant 893/15.

References and links

1. W. F. Krupke, “Diode pumped alkali lasers (DPALs) — A review (rev1),” J. Progr. Quantum Electron. 36(1), 4–28 (2012). [CrossRef]  

2. B. V. Zhdanov and R. J. Knize, “Review of alkali laser research and development,” Opt. Eng. 52(2), 021010 (2012). [CrossRef]  

3. A. V. Bogachev, S. G. Garanin, A. M. Dudov, V. A. Eroshenko, S. M. Kulikov, G. T. Mikaelian, V. A. Panarin, V. O. Pautov, A. V. Rus, and S. A. Sukharev, “Diode-pumped caesium vapour laser with closed-cycle laser-active medium circulation,” J. Progr. Quantum Electron. 42(2), 95–98 (2012). [CrossRef]  

4. B. V. Zhdanov, M. D. Rotondaro, M. K. Shaffer, and R. J. Knize, “Potassium diode pumped alkali laser demonstration using a closed cycle flowing system,” Opt. Commun. 354, 256–258 (2015). [CrossRef]  

5. B. D. Barmashenko and S. Rosenwaks, “Feasibility of supersonic diode pumped alkali lasers: model calculations,” Appl. Phys. Lett. 102(14), 141108 (2013). [CrossRef]  

6. E. Yacoby, K. Waichman, O. Sadot, B. D. Barmashenko, and S. Rosenwaks, “Modeling of supersonic diode pumped alkali lasers,” J. Opt. Soc. Am. B 32(9), 1824–1833 (2015). [CrossRef]  

7. Z. Yang, H. Wang, Q. Lu, Y. Li, W. Hua, X. Xu, and J. Chen, “Modeling, numerical approach, and power scaling of alkali vapor lasers in side-pumped configuration with flowing medium,” J. Opt. Soc. Am. B 28(6), 1353–1364 (2011). [CrossRef]  

8. B. D. Barmashenko and S. Rosenwaks, “Modeling of flowing gas diode pumped alkali lasers: dependence of the operation on the gas velocity and on the nature of the buffer gas,” Opt. Lett. 37(17), 3615–3617 (2012). [CrossRef]   [PubMed]  

9. B. D. Barmashenko and S. Rosenwaks, “Detailed analysis of kinetic and fluid dynamic processes in diode-pumped alkali lasers,” J. Opt. Soc. Am. B 30(5), 1118–1126 (2013). [CrossRef]  

10. S. Rosenwaks, B. D. Barmashenko, and K. Waichman, “What can we gain from supersonic operation of diode pumped alkali lasers: model calculations,” Proc. SPIE 8898, 88980X (2013). [CrossRef]  

11. B. D. Barmashenko, S. Rosenwaks, and K. Waichman, “Model calculations of kinetic and fluid dynamic processes in diode pumped alkali lasers,” Proc. SPIE 8898, 88980W (2013). [CrossRef]  

12. B. D. Barmashenko, S. Rosenwaks, and K. Waichman, “Kinetic and fluid dynamic processes in diode pumped alkali lasers: semi-analytical and 2D and 3D CFD modeling,” Proc. SPIE 8962, 86920C (2014).

13. S. Rosenwaks, B. D. Barmashenko, and K. Waichman, “Semi-analytical and 3D CFD DPAL modeling: Feasibility of supersonic operation,” Proc. SPIE 8962, 896209 (2014). [CrossRef]  

14. K. Waichman, B. D. Barmashenko, and S. Rosenwaks, “CFD DPAL modeling for various schemes of flow configurations,” Proc. SPIE 9251, 92510U (2014).

15. K. Waichman, B. D. Barmashenko, and S. Rosenwaks, “Computational fluid dynamics modeling of subsonic flowing-gas diode pumped alkali lasers: comparison with semi-analytical model calculations and with experimental results,” J. Opt. Soc. Am. B 31(11), 2628–2637 (2014). [CrossRef]  

16. A. H. Shapiro, The Dynamics and Thermodynamics of Compressible Fluid Flow, Vol. 1 (Ronald, 1953), pp. 73–152.

17. B. D. Barmashenko and S. Rosenwaks, “Analysis of the optical extraction efficiency in gas-flow lasers with different types of resonator,” Appl. Opt. 35(36), 7091–7101 (1996). [CrossRef]   [PubMed]  

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 Schematics of the transonic (a), supersonic (b) and subsonic (c) devices with transverse pumping by rectangular pump beam of 5 x 2 cm2 aperture. The flow is perpendicular to the laser and pump beams propagation axes. Note that different buffer gases are used for Cs and K.
Fig. 2
Fig. 2 The mesh for the rectangular transverse-pumping beam, consists of nx x ny x nz hexahedral cells (nx = 30, ny = 80, nz = 80).
Fig. 3
Fig. 3 Dependence of Plase on Pp for subsonic (M ~0.2) transonic (M ~0.9) and supersonic (M ~2.7) Cs DPALs with CH4 as a buffer gas, and with optimal parameters summarized in Table 1.
Fig. 4
Fig. 4 Dependence of Plase on Pp for subsonic (M ~0.2) transonic (M ~0.9) and supersonic (M ~2.4) К DPALs with He as a buffer gas, and with optimal parameters summarized in Table 2.

Tables (2)

Tables Icon

Table 1 Optimal parameters of different Cs DPALs with CH4 as a buffer gas, transversely pumped with 5 x 2 cm2 rectangular pump beam, with reflector r1 = 0.95, output coupler r2 = 0.35 and windows transmission t = 0.98. For transonic flow pstag = 10 atm, Tstag = 425K and for supersonic pstag = 70 atm, Tstag = 493 K.

Tables Icon

Table 2 Optimal parameters of different K DPALs with He as a buffer gas, transversely pumped with 5 x 2 cm2 rectangular pump beam, with reflector r1 = 0.95, output coupler r2 = 0.35 and windows transmission t = 0.98; the pump bandwidth Δ v p = 10 GHz. For transonic flow pstag = 11 atm, Tstag = 485K and for supersonic pstag = 90 atm, Tstag = 586 K.

Equations (15)

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

T i T stag = ( p i p stag ) γ1 γ = ( n X,in n X,stag ) γ1 = ( 1+ γ1 2 M i 2 ) 1 ,
n sat = p sat k B T stag 10 a b T stag k B T stag .
n X,stag = n X,i ( 1+ γ1 2 M i 2 ) 1 γ1
I p ± (x,y,z,ν) y =± g 31 (x,y,z,ν) I p ± (x,y,z,ν),
g 31 ( x,y,z,ν )= σ 31 ( x,y,z,ν )[ n 3 ( x,y,z )2 n 1 ( x,y,z ) ],
  σ 31 (x,y,z,ν)= σ 31 (x,y,z, ν p ) f 31 (ν)
  σ 31 (x,y,z,ν)= σ 31,at Δ ν 31,n Δ ν 31 (x,y,z) ,
f 31 (ν) = (Δ ν 31 /2) 2 /[ ( ν   ν p ) 2 + (Δ ν 31 /2) 2 ]     
I p ± ( ν )=  P p WL t f p ( ν ),
f p (ν)= ( 4ln2 πΔ ν p 2 ) 1/2 exp[ 4ln2 ( ν ν p Δ ν p ) 2 ]
I l ± ( x,y,z ) z =± g 21 ( x,y,z ) I l ± ( x,y,z ),
g 21 ( x,y,z )= σ 21 ( x,y,z, ν l   )[ n 2 ( x,y,z ) n 1 ( x,y,z ) ],
σ 21 (x,y,z, ν l )= σ 21,at Δ ν 21,n Δ ν 21 (x,y,z)
P lase = 0<x<L,0<y<H I lase ( x,y )dxdy .
n 3 n 2 =2exp( ΔE k B T ).
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.