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

Pulse train characteristics of a passively Q-switched microchip laser

Open Access Open Access

Abstract

A study of passively Q-switched microchip laser pulse trains yields approximate, yet reliable, formulae for the peak power, pulse energy, half-width, period, and the pulse shape in time. The pulse gain differential equation is made integrable by assuming that the laser absorption cross sections for the gain and saturable absorber are equal. We compare our predictions with an experiment which uses Nd:YAG as a gain medium and Cr:YAG as a saturable absorber. The agreement between theory and experiment for the period, pulse width, and the pulse energy is within 10%.

©1999 Optical Society of America

1. Introduction

During the past few years there has been an ever increasing interest in experimental studies of passively Q-switched(PQS) microchip lasers. A detail review of devices using solid state saturable absorbers is given in [1], [2]. However, the same trend is not evident in theoretical modeling[3, 4, 5, 6] studies. Only recently[7] has there been a comprehensive theoretical and numerical investigation of the pulse dynamics. This study included the limit-cycle analysis and presented a successful numerical analysis of the differential equations in the numerically challenging region where the pulse separation is six orders-of-magnitude larger than the pulse width. However, one area not thoroughly addressed was the pulse shape. In the following we derive simple design formulae for the peak power, pulse width, repetition rate, pulse energy and the pulse shape for microchip lasers which are derived with the experimentalists in mind. These approximate solutions are anchored to our previous simulations[7] which agree with experiment[8]. Finally, since our solutions are approximate we also discuss their limitations.

Initial studies of Q-switching were concerned with active[9] and then passive switching[10, 11, 12] of single pulses. These studies described numerically the pulse shape for various losses, and cross sections. A later paper[13] devoted a small section to pulse trains based on single pulse results. And, finally, a very recent paper[14] divided the pulse into four regions and deduced some design equations. Their development is for a semiconductor saturable absorber which has a decay rate much faster than the gain medium. Also, they restrict the saturable absorber unsaturated losses to be much less than the outcoupling losses, and finally they consider operation just above threshold. These assumptions lead to a symmetric pulse shape, for example. Subsequently, we showed[15] that their major results are contained within our earlier work[7]. In the following we relax these restrictions and develop design equations which include solid state saturable absorbers as well. Generally, the limit-cycle solution for passive Q-switching is described by a set of five coupled equations for the period, and the initial and final values of the laser gain and saturable absorber gain[7]. Our problem here is to reduce these equations to a reliable and managable form for solid state abosrbers.

The following development relies on our recent work[7]. Here, however, we develop the gain quick-pulse differential equation in more detail. In fact we are able to integrate it after introducing several reasonable approximations which results in the gain as an implicit function of time. The integration of the gain equation depends on the initial laser gain before the pulse. This, as just mentioned, is coupled to the period and the final gain after the pulse along with the initial and final saturable absorber gains. In this paper we show that the conditions under which these equations reduce to a single equation correspond to solid state saturable absorber parameters. We compare the approximate solution to the exact numerical simulations.

We close by restating the difference between previous singel pulse work in lasers and our studies. Pulsing in lasers is intrinsically a problem in laser instability and is physically and mathematically quite different from the single pulse situation. This difference is partly embodied in the period equation which is critical in determining the pulse characteristics.

2. Theory

We treat the gain medium as an ideal four-level system where the decay from the lower laser level to the ground state is nonradiative and very fast[7, 16, 17]. This process is described by the dimensionless gain G≡2LgγeN 2. N 2 is the upper lasing level concentration, and N 0 is the ground state concentration with N 0+N 2=NT where NT is the doping concentration. The gain length is Lg and the laser effective emission cross section is γe. On the other hand, we treat the saturable absorber as a two-level system described by the dimensionless gain Gs≡2Ls(σeN 2-σaN 1) where N 2 is the upper level concentration and N 1+N 2=N 0, where the total doping concentration of the saturable absorber is N 0. The saturable absorber length is Ls and the laser flux is coupled through the effective emission cross section σe and the effective absorption cross section σa. Finally, the gain decay rate is Ag=1/τg and the saturable absorber decay rate is As=1/τs. The pump effective absorption cross section is γ′a.

Using these notations, the upper level rate equation for the gain medium is

N˙2=N˙0=PγaN0RγeN2N2τg

which yields

G˙=1τg[2LgNTPΓτg(1+γaτgP+γeτgR)G]

where Γ=γ′aγe, R(t) is the laser flux, and P is the cw pump flux. The saturable absorber is a two-level atom with the rate equation for the upper lasing level given by 2=- 1=R(γaN 1-γeN 2)-N 2=τg. Thus, the differential equation for the dimensionless saturable absorber gain Gs is

G˙s=1τs[2Ls𝒩0σa(1+σ+τsR)Gs]

where σ +=σe+σa. Accompanying eqs. (2,3) is the differential equation for the sum of the forward and reverse laser fluxes R(t). In the mean field approximation this sum obeys

R˙=1τc[G+Gs(ln(1r)+L)]R.

The cavity round trip time is τc=2Lc/v where v=c/n, c is the speed of light in vacuum, n is the index of refraction, and Lc=Lg+Ls. The laser outcoupling reflectivity is r and all other losses are included in L. Using the above equations we form the following three dimensionless differential equations

dIds=I(1+AD+A-D-),dDds=γ(1D(1+I)),dD-ds=γ-(1D-(1+αI)),

where I, D, and are the dimensionless laser intensity, dimensionless gain, and the dimensionless saturable absorber inversion defined by

D(1+γaτgP)G2LgNTPΓτg,D-Gs2Ls𝒩0σa,IγeτgR(1+γ'aτgP).

The parameters in eq. (5) are related to the physical constants through

A=2LgNT(PΓτgγa)αL(1+γaτgP),A-=2Lsσa𝒩0αL,γ=τc(1+γaτgP)τgαL,γ-=τcτsαL,
α=σ+τsγaτg(1+γaτgP),αL=ln(1r)+L,s=αLtτc.

During the pulse when, I>>1, the three above normalized differential equations can be reduced to a single differential equation for the gain[7]. This is accomplished by dividing the second equation in eq.(5) by the third equation and integrating. This step allows elimination of when the first equation in eq. (5) is divided by the second equation. Integrating this last division gives

γI=ADb(1η)+lnη+γA-D-bαγ-[1ηm],

and thus the dD=ds equation becomes

dηds=η(ADb(1η)+lnη+γA-D-bαγ-[1ηm]),whereη=DDb,m=(αγ-γ).

Db is the dimensionless gain inversion just before the pulse starts; we discuss this in more detail later.

Our objective is to develop approximate, yet reliable, analytic solutions for the microchip laser pulse intensity. To this end, we rely on the results of our numerical simulations of eqs. (9,10). These show that if m is set equal to unity then the pulse shape is very close to the actual pulse shape obtained when microchip constants are used. Note that for the microchip experiment the exponent m=αγ̄/γ=σ +/γa=3.2. The choice of m=1 allows us to solve eq. (10) analytically. The restriction m=1 requires the laser absorption cross sections of the gain to equal that of the saturable abosrber. This constraint, however, gives good results since the pulse shape is not strongly dependent on m, rather the shape depends on other parameters as we discuss near the end of the next section. Figures (1b,c) show the pulses for the m=3.2 and the m=1 solutions, respectively. The approximate m=1 pulse (curve c) turns on slightly sooner and is slightly higher than the exact solution (curve b). This small shift is inconsequential since the period is roughly six orders-of-magnitude greater than the pulse width. Simulations of eqs. (5) further show that at the initiation of the pulse the saturable absorber is bleached, that is b=-1, and that the gain is extracted, i. e. Da<<Db where Da is the gain after the pulse.

 figure: Figure 1.

Figure 1. Simulated pulse shapes as a function of time. (a), the three differential equations, eq. (5); (b), the gain differential equation for the actual microchip pa- rameters m=3:2, eq.(10); (c), gain differential equation for m=1, eq. (10); (d), the implicit solution, eq. (19).

Download Full Size | PDF

With these results in mind we set m=1 and proceed to derive various approximations to the exact pulse shape including the peak power and the FWHM. For m=1 the normalized gain differential equation, eq. (10), can be written as

dηdsη(lnη+ADbβ(1η)),whereβ=1A-γαγ-ADb.

We solve the above equation using integration by parts. The integral generated in this process can be approximated with little error by noting that most of the contribution comes from the region near η=1 where ln η≈2(η-1)=(η+1). Thus, continuing with the algebra, we obtain the implicit solution

Δsln[lnη+C(1η)]CC1ln(1η)+1C1ln(Cη+C2),

accompanied by the intensity equation

γIlnη+C(1η)

where C=ADbβ. The pulse shape embodied in eq. (12,13) is very close to the full simulations of eqs. (5) and in a moment we will make this comparison. But before doing so we derive equations for the peak power Ip and the pulse width (FWHM). At the pulse peak the intensity differential equation gives -1+ADp+ĀD̄p=0. Also, at the peak our numerical simulations show that the saturable absorber is transparent, i. e. p≈0. Thus, Dp≈1/A. Inserting this back into eqs. (12,13) gives the peak intensity Ip as

γIpln(1ADb)+C(11ADb).

Consequently, the values for η at the halfwidth are obtained from

γIt,rlnηt,r+C(1ηt,r)=γIp2.

This equation has two roots which determine the half-width inversions during the rise of the pulse, ηr, and in the pulse tail, ηt. In the tail region ηt is small and in the rise region ηr is near one and thus eq. (15) yields the approximate equations

ηtexp(γIp2C),andηr1+γIp2C1C.

The half-width is then determined from eq. (12) by inserting ηr,t and forming the FWHM ΔsFWstsr. Completing this task gives our final result

ΔsFWCC1ln[2(C1)γIp]+1C1ln[2(C1)2+CγIp2(C1)(C2)].

As we will show the FWHM given by eq. (17) is within 3% of the microchip laser experimental values[8].

Even though the implicit equations, eqs. (12,13), are approximate they are never-the-less a good representation of the exact pulse shape. However, our development would be more informative if we derive an approximate explicit form revealing the growth and decay time scales. To this end we find a piecewise solution. In the tail region where η<<1 eqs. (12,13) show that the intensity It is given by

γItexp(Δs).

This decay is relatively slow and depends only on the cavity decay rate. Its typical time scale is s-s*=O(1); s* is the integration constant and is unimportant in this development. In the pulse rise region substitution of the expansion ln η≈2(1-η)/(1+η) into eqs. (9,10), for m=1, followed by integration gives

γIr=Cexp((C2)Δs)C2+exp((C2)Δs)+ln(C2C2+exp((C2)Δs)).

Thus, the initiation of the pulse is fast and is clearly controlled by the pump in a complex manner. Its typical time scale is s-s*=O(1/(ADbβ))<<1.

Next we briefly discuss the period T and initial inversion Db in PQS lasers; for a more detailed discussion see reference [5]. Between pulses I is much less than one, consequently the gain equations in eq. (5) integrate into exponentials. Furthermore our simulations assure that after the pulse Da;a<<1. Also, the fact that the period is much longer than the saturable absorber decay time implies that before the pulse b-1, as stated before. These assumptions lead to the inversion equation D=1-exp(-γs). After this equation is inserted into the flux equation in eq. (5) and integrated the resultant exponential function for the intensity I is small until s reached S where S satisfies

(1+AA-)XA(1exp(X))+A-γγ-=0,whereDb=1exp(X),

where X=γS. When eq. (20) is satisfied a rapid pulse growth is initiated. This defines the period S. As we will see the value of S agrees with the experiment[8] to within a few percent.

In summary, modeling a particular experiment is straight forward. After all the physical constants and the pump flux have been identified, one forms parameters A, Ā, , and γ̄ using eqs. (6, 7, 8). Next, eq. (20) is solved for X from which the period T, and initial population inverison Db can be calculated through T=Xτc/αL and Db=1- exp(-X), respectively. After Db is determined the peak intensity is obtained using eq. (14) and the half-width using eq. (17). Finally, the pulse energy ε can be found through[7]

εt(Area)hνγe[ADbA-γαγ-]

where t is the outcoupling intensity transmission, t+r=1, and hv is the laser photon energy. Of course, all this can be done as a function of the pump power. Finally, the pulse profile can be generated by solving one of the differential equations eq.(10) or eq.(11), or by plotting the implicit solution as a simple loop over η, see eq. (12). Alternatively, one could also use the explicit solution given by eq. (19).

3. Example

Now we model the microchip laser experiment[8] which uses Nd:YAG as a gain medium of length Lg=.05cm butt coupled to a Cr+4:YAG saturable absorber of length :025cm. The outcoupling reflectivity is given as r=94%. The Nd doping is at 1.6at.%. The absorption coefficient[8] of Cr is estimated at 6cm-1 for a wavelength of 1.064µm. The experimental report is brief but lists a laser repetition rate of 6kHz, pulse width of 330ps, pulse energy of 11µJ, peak pulse power of 27kW at a peak intensity of 180MW/cm2. This behavior is observed for a pump power of 1.2W with a threshold at .8W. The laser emits at a wavelength of λl=1.06µm and has an estimated beam waist of 70µm. The pump with a wavelength of λp=808nm is double passed since it reflects from the Nd-Cr interface.

Prior to discussing the predictions of this paper we summarize our previous modeling[7]. The full simulation used a patch technique[7] applied during the inter-pulse period when the laser energy is negligible and, consequently, the two inversion equations are easily integrable. In this work we estimated the values of the dimensionless parameters [7] as

γ=1.75×106,γ-=6.35×105,A-=3.96,α=0.0852.

Moreover, for a pump power of 1.2W, A=10.48 and the threshold (Ath≡1+Ā)=4.96. Also, we found b=-1, and a=10-6, Da=10-4, and Db=.759. Furthermore these simulation of eqs. (5) gave a period of 5.9kHz, a peak power of 29.8kW, the pulse width as 343psec, and a pulse energy of 11.9µJ. These values are very close to the above mentioned experimental values. Figure 2 shows the simulated pulse shape, the solid curve, compared to the experimental results shown as the dashed curve. Note that since our simulation does not include noise, the intensity drops to zero much more rapidly than in the experiment[8]. However, the simulation does capture the pulse rise and decay.

 figure: Figure 2.

Figure 2. Pulse shape as a function of time in 200psec/division. Experimental, dashed curve, simulation, solid curve.

Download Full Size | PDF

Now we return to fig. (1) which summarizes the major results of this paper. For the sake of reference we repeat the full simulation[7] shown in fig. (2) as curve (a). As mentioned earlier this simulation gives a value of Db=.759. At this point we employ our approximate equations. First, inserting the above values for A, Ā, γ, and γ̄ into eq. (20) gives a period of 5.64kHz, which is within 7% of the experiment. This value of the period leads to a value of Db equal to .7584. We use this value as the initial condition for solving the inversion differential equation, eqs. (9,10) for the two values of m=αγ̄/γ=3.2 and m=1 which yields curves (b) and (c), respectively. The final curve (d) is our implicit solution [see eqs. (12,13)]. These four simulated pulse shapes are very close to one another. The two rigorous solutions, curve (a) and curve (b), are almost identical. The m=1 differential equation, eqs. (9,10), over estimates the peak and the width by about 7%. Our implicit solution, eqs. (12, 13), closely replicates the simulated solutions in curves (a) and (b). In fact, further simulations show that the pulse shape remains within 10% of the experimental values for 1<m<4.

The merit of our analysis lies in the application of our approximate equations. Now that we have determined the period and Db, eq. (14) gives a peak intensity of Ip=206MW/cm2, eq. (17) gives the half width of ΔsFW=321psec, and finally eq. (21) gives the energy ε=12.1µJ. The peak intensity matches the experiment to within less that 15% while the others agree to less than 10%.

Further information about the pulse shape is shown in fig. (3). This figure shows that the pulse rise time (1=(C-2)), obtained from eqs. (18,19), is a strong function of pump power. However, as we have mentioned the pulse decay depends only on the cavity decay time which is 216psec. Thus, as the pumping is changed the only pulse shape changes occurs in the early part of the pulse. Note that for large pumping eq. (17) shows that the minimum pulse width is governed by the cavity and is given by (τc/αL) ln(2). This value, of course, is not a practical limit since it corresponds to a delta function like pulse growth.

 figure: Figure 3.

Figure 3. The pulse rise time as a function of pump power.

Download Full Size | PDF

We conclude this comparison by mentioning that further numerical simulations show the above approximations are good for as much as 10 times above threshold before the error between the simulations and the approximations exceeds 20%. Also. these simulations show that the pulse shape, and its associated parameters such as the pulse energy, are mild function of m=σ +/γ. Specifically for.5<m<5 the pulse energy agrees with the m=3.2 solution to within about 15%. However, the pulse is strongly dependent on τg, τs, the pump small signal gain gg=2γNTLg, and the laser small signal gain gs=2σaN0Ls in the saturable absorber. Specifically eq. (21) shows that ε∝ (ggPγeτg-gsτs/τg)/αL.

Finally, we turn to a discussion of a recent experiment[14] which uses either Nd:YVO4 or Nd:LSB as a gain medium and a SESAM (seminconductor saturable absorber mirror). When we compare their pulse differential equations with our normalized pulse form, eqs.(5), we find that: γ̄=9.3×10-2, γ=3.7×10-7, Ā=.36, A=4.08, α=4.×10-3, and that m is large with a value of m=1000. After these values are inserted in the exact gain differental equation eq.(10), for b=-1, and in the m=1 differential equation eq. (11) we find that the two pulse shapes are virtually identical. The reason that the m=1 solution works so well is because the [1-ηm] term is multipled by 1/m, or β≈1. An absolute comparison with their experiment[14] is difficult since we must know cross sections, doping levels, losses, etc., and these are not readly available. Finally, as we mentioned in the introduction, we derived their formulas[15] published subsequent to their report[14].

4. Summary

We have developed accessible approximate analytic equations which characterize the pulse shape in microchip lasers. We have shown that it is straight forward to calculate the pulse period, peak power, pulse width, and pulse energy along with the pulse profile.

We would like to acknowledge Thomas Erneux at the Université Libre de Bruxelles, Optique Nonlinéaire Théorique, Campus Plaine, C. P. 231, 1050 Bruxelles, Belgium for his suggestions.

References

1. John J. Zayhowski, “Passively Q-switched Microchip Lasers and Applications,”The Review of Laser Engineering,” 26, 841–846 (1998).

2. John J. Zayhowski, “Microchip Lasers.” Optical Materials 11, 255–267 (1999). [CrossRef]  

3. J. J. Zayhowski and P. L. Kelly, “Optimization of Q-switched lasers,” IEEE J. Quantum. Electron. 27, 2220–2225 (1991). [CrossRef]  

4. J. J. Degnan, “Theory of the optimally coupled Q-switched laser,” IEEE J. Quantum Electronics 25, 214–220, (1989). [CrossRef]  

5. X. Zhang, S Zhao, Q. Wang, Q. Zhang, L. Sun, and S. Zhang, “Optimization of Cr4+ doped saturable absorber Q-switched lasers,” IEEE J. Quantum Electronics 33, 2286–2294, (1997). [CrossRef]  

6. A. Agnesi, S Dell’Acqua, C. Morello, G Piccino, G. C. Reali, and Z. Sun, “Diode-pumped Neodymium laser repetitively Q-switched by Cr4+:YAG solid-state saturable absorbers,” IEEE J. Selected Topics in Quantum Electronics 1, 45–52, (1997). [CrossRef]  

7. P. Peterson, A. Gavrielides, M.P. Sharma, and T. Erneux, “Dynamics of passively Q-switched microchip lasers,” IEEE J. Quant.Electr. 35, 1–10, (1999). [CrossRef]  

8. J. J. Zayhowski and C. Dill III, “Diode-pumped passively Q-switched picosecond microchip laser ’,” Opt. Lett. 19, 1427 (1994). [CrossRef]   [PubMed]  

9. W. G. Wegnar and B. A. Lengel, “Evolution of the giant pulse in a laser,” J. Appl. Phys. 42, 2040–2046 (1963). [CrossRef]  

10. A. Szabo and R. A. Stein, “Theory of giant pulsing by a saturable absorber,” J. Appl. Phys. 36, 1562–1566 (1965). [CrossRef]  

11. L. E. Erickson and A Szabo “Effects of saturable absorber lifetime on the performance of giant-pulse lasers,” J. Appl. Phys. 37, 4953–4961 (1966). [CrossRef]  

12. L. E. Erickson and A Szabo “Behavior of saturable absorber giant-pulse lasers in the limit of large absorber cross section,” J. Appl. Phys. 38, 2540–2542 (1967). [CrossRef]  

13. J. J. Degnan,“Optimization of passively Q-switched lasers, ”IEEE J. Quantum Electron. 31, 1890–1902 (1995). [CrossRef]  

14. G. J. Spuhler, R. Paschotta, R. Fluck, B. Braun, M. Moser, G. Zhang, E. Gini, and U. Keller, “Experimentally confirmed design guidelines for passively Q-switched microchip lasers using semiconductor saturable absorbers,” J. Opt. Soc. 16, 376–388 (1999). [CrossRef]  

15. T. Erneux, P. Peterson, and A. Gavrielides, “The pulse shape of a passively Q-switched microchip laser,” accepted Europ. J. Physics (1999).

16. Peter W. Milonni and Joseph H. Eberly, Lasers (John Wiley and Sons, New York, 1998).

17. Walter Koechner, Solid-State Laser Engineering (Springer, New York, 1992).

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 (3)

Figure 1.
Figure 1. Simulated pulse shapes as a function of time. (a), the three differential equations, eq. (5); (b), the gain differential equation for the actual microchip pa- rameters m=3:2, eq.(10); (c), gain differential equation for m=1, eq. (10); (d), the implicit solution, eq. (19).
Figure 2.
Figure 2. Pulse shape as a function of time in 200psec/division. Experimental, dashed curve, simulation, solid curve.
Figure 3.
Figure 3. The pulse rise time as a function of pump power.

Equations (22)

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

N ˙ 2 = N ˙ 0 = P γ a N 0 R γ e N 2 N 2 τ g
G ˙ = 1 τ g [ 2 L g N T P Γ τ g ( 1 + γ a τ g P + γ e τ g R ) G ]
G ˙ s = 1 τ s [ 2 L s 𝒩 0 σ a ( 1 + σ + τ s R ) G s ]
R ˙ = 1 τ c [ G + G s ( ln ( 1 r ) + L ) ] R .
dI ds = I ( 1 + AD + A - D - ) , dD ds = γ ( 1 D ( 1 + I ) ) , d D - ds = γ - ( 1 D - ( 1 + α I ) ) ,
D ( 1 + γ a τ g P ) G 2 L g N T P Γ τ g , D - G s 2 L s 𝒩 0 σ a , I γ e τ g R ( 1 + γ ' a τ g P ) .
A = 2 L g N T ( P Γ τ g γ a ) α L ( 1 + γ a τ g P ) , A - = 2 L s σ a 𝒩 0 α L , γ = τ c ( 1 + γ a τ g P ) τ g α L , γ - = τ c τ s α L ,
α = σ + τ s γ a τ g ( 1 + γ a τ g P ) , α L = ln ( 1 r ) + L , s = α L t τ c .
γ I = A D b ( 1 η ) + ln η + γ A - D - b α γ - [ 1 η m ] ,
d η ds = η ( A D b ( 1 η ) + ln η + γ A - D - b α γ - [ 1 η m ] ) , where η = D D b , m = ( α γ - γ ) .
d η ds η ( ln η + A D b β ( 1 η ) ) , where β = 1 A - γ α γ - A D b .
Δ s ln [ ln η + C ( 1 η ) ] C C 1 ln ( 1 η ) + 1 C 1 ln ( C η + C 2 ) ,
γ I ln η + C ( 1 η )
γ I p ln ( 1 A D b ) + C ( 1 1 A D b ) .
γ I t , r ln η t , r + C ( 1 η t , r ) = γ I p 2 .
η t exp ( γ I p 2 C ) , and η r 1 + γ I p 2 C 1 C .
Δ s FW C C 1 ln [ 2 ( C 1 ) γ I p ] + 1 C 1 ln [ 2 ( C 1 ) 2 + C γ I p 2 ( C 1 ) ( C 2 ) ] .
γ I t exp ( Δ s ) .
γ I r = C exp ( ( C 2 ) Δ s ) C 2 + exp ( ( C 2 ) Δ s ) + ln ( C 2 C 2 + exp ( ( C 2 ) Δ s ) ) .
( 1 + A A - ) X A ( 1 exp ( X ) ) + A - γ γ - = 0 , where D b = 1 exp ( X ) ,
ε t ( Area ) h ν γ e [ A D b A - γ α γ - ]
γ = 1.75 × 10 6 , γ - = 6.35 × 10 5 , A - = 3.96 , α = 0.0852 .
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.