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

Degree of phase-space separability of statistical pulses

Open Access Open Access

Abstract

We introduce the concept of phase-space separability degree of statistical pulses and show how it can be determined using a bi-orthogonal decomposition of the pulse Wigner distribution. We present explicit analytical results for the case of chirped Gaussian Schell-model pulses. We also demonstrate that chirping of the pulsed source serves as a powerful tool to control coherence and phase-space separability of statistical pulses.

© 2012 Optical Society of America

1. Introduction: Wigner function description of statistical pulses

The acute recent interest in femtosecond pulses and pulse sources [1] has triggered resurgence of activity in the field of ultrafast statistical optics [211]. The latter should come as no surprise as statistical properties of ultrashort pulses impose ultimate limits on the performance and accuracy of the state-of-art fiber-optical communication systems, for instance; not to mention that spontaneous emission noise degrades the output coherence of optical amplifiers [12]. To date then, there has been extensive research done on modeling statistical properties of realistic sources of ultrashort pulses [2, 3]. The fundamental issues of defining and measuring the spectrum of statistical pulses [4] and formulating coherence theory of periodic statistical pulse trains [5, 6] have been addressed. Due to its relevance for fiber optical communications, the propagation of ultrashort statistical pulses in linear [13,14] and nonlinear [15] dispersive media has also been explored. Further, several approaches have been recently advanced to synthesize novel partially coherent pulses from uncorrelated–or partially correlated–superpositions of elementary pulses in time [7] and frequency [8]. In addition, several phase-space approaches to partially coherent pulse representation were discussed in the literature [9,10]. Lately, a general phase-space approach has been put forward to describe partially coherent pulse synthesis from complex Gaussian pulses [11]. A key feature of the complex Gaussian representation is its versatility: It can be used to either generate new partially coherent pulses or represent the ones with known cross-correlation functions in terms of statistically uncorrelated Gaussian pulses.

In this work, we show that a Wigner distribution based phase-space representation for statistical pulses provides a natural context to define a measure of their phase-space separability. Next, we show how the introduced degree of phase-space separability can be determined using a bi-orthogonal decomposition of the Wigner distribution of the pulse. We then discuss the way the new measure changes on chirped pulse propagation in linear dispersive media. Our results are directly relevant for temporal imaging with ultrashort pulses which involves temporal lenses, chirping the pulses, and dispersive delay lines, e. g., linear optical fibers [16]. To make our results more instructive, we specialize to a representative case of chirped Gaussian Schell-model pulses for which closed-form analytical results can be obtained. It follows from our results that chirping ultrashort partially coherent pulses can provide a potent tool for controlling their degrees of coherence and phase-space separability.

To set the stage, we consider an ensemble of random pulses {E(t)} and decompose the electric field E(t) into a slowly-varying envelope U(t) and a carrier wave [17] such that

E(t)=U(t)eiωct,
where ωc is a deterministic carrier frequency of the pulse. The second-order statistical properties of the ensemble {U(t)} are specified by the cross-correlation function
Γ(t1,t2)=U*(t1)U(t2),
where the angle brackets denote ensemble averaging. The Wigner distribution (WD) of the ensemble is then defined as
𝒲(t,ω)=dτΓ(tτ/2,t+τ/2)eiωτ,
where we introduced the variables t = (t1 + t2)/2 and τ = t1t2. The intensity I(t) and spectrum S(ω) of the pulse are determined by the appropriate marginals of the Wigner distribution viz.,
I(t)=dω𝒲(t,ω);S(ω)=dt𝒲(t,ω).
Many phase space characteristics of the pulses, such as the position of the pulse center (in time), central frequency of the envelope, rms temporal and spectral widths etc., can be determined as the corresponding moments of I or S. Notice also that it follows at once from Eqs. (2) and (3) that fully as well as partially temporarily and/or spectrally coherent pulses can be treated in the same way using the Wigner distribution, the fully coherent case being just the limiting situation when the cross-correlation function factorizes.

In this work, we will examine the WD evolution as pulses propagate in transparent homogeneous linear dispersive media. To this end, recall that the pulse envelope in such media obeys the paraxial wave equation [18]

izU12β2ss2U=0,
where we introduced the retarded time s = tβ1z, β1 being the inverse group velocity, and β2 is the group-velocity dispersion (GVD) coefficient. It can be inferred at once from Eqs. (2) and (5) that Γ obeys the evolution equation in the form
izΓ+12β2(s1s12s2s22)Γ=0.
Thus, the WD propagation is governed by the first-order equation
z𝒲+β2T𝒲=0,
where the variables
T=s1+s22,andτ=s1s2,
were introduced. Solving Eq. (7) on characteristics, we obtain
𝒲(T,ω;z)=𝒲0(Tβ2ω,ω).
In other words, as a statistical pulse propagates in a linear dispersive medium such as an optical fiber, say, its Wigner distribution in any transverse plane z = const can be related to its form in the source plane, 𝒳0(t,ω), by a rather simple equation. Equation (9) is a temporal analog of the well-known transformation law describing passage of partially coherent beams through first-order optical systems [19, 20] and is a direct consequence of space-time analogy [2123] between beam and pulse propagation in free space and linear dispersive media, respectively.

2. Quantifying phase-space separability of statistical pulses

The Wigner distribution contains all available information about second-order statistics of the pulses. This information can be revealed by directly measuring the set of phase-space projections of WD known as a Radon-Wigner transform [24]. In general, however, such measurements are rather involved. The situation drastically simplifies for WDs in the separable form such that

𝒲(t,ω)I(t)S(ω).
The separability of Eq. (10) implies that the second-order statistics of the system can be recovered by separately measuring the intensity profile and spectrum of the pulse. This observation prompts the question: Are there any realistic pulses with a separable WD? A related fundamental issue is: Given an arbitrary statistical ensemble of pulses, how can one quantitatively describe phase-space separability of its WDs? And to follow up on this: How can one control such separability, if at all?

To address the first question, it is sufficient to recall that a common technique to generate statistical pulses involves temporal chopping of statistically stationary light fields [3]. We can consider, for instance, chopping a Gaussian correlated statistically stationary field with a Gaussian temporal modulation function. This procedure yields the so-called Gaussian Schell-model (GSM) source [2] with the correlation function that can be transformed to

ΓGSM(t1,t2)=I(t1+t22)g(t1t2),
where both I and g are Gaussians. It follows at once from Eq. (3) and (11) that the WD of a GSM source has a separable form of Eq. (10). Thus, realistic statistical pulses belonging to a wide GSM class have separable WDs. Moreover, the cross-correlation function of any quasi-stationary source can be well approximated as
Γqs(t1,t2)I(t1+t22)γ(t1t2),
where I(t) is a “slowly-varying” intensity profile and γ(t) is a “fast” temporal degree of coherence. The WD of such sources are approximately separable as well.

Unfortunately, even if the WD of a source is separable, the WD of generated pulses need not be so as is seen from Eq. (9). In general, the pulse propagation even in a linear dispersive medium couples phase-space variables, providing additional impetus for our quest for a phase-space separability measure. To address this issue, we propose to expand the pulse WD in any transverse plane into a bi-orthogonal series as

𝒲(T,ω;z)=nλn(z)χn(T,z)ϕn(ω,z),
where the eigenvalues {λn} and eigenfunctions, {χn} and {ϕn}, are real due to reality of the WD. It is known [25] that the two sets of eigenfunctions obey the Fredholm integral equations which, in our case, take the form
λn(z)ϕn(ω,z)=dT𝒲(T,ω;z)ϕn(T,z),
and
λn(z)χn(T,z)=dω𝒲(T,ω;z)χn(ω,z).
The suitably normalized eigenfunctions form an orthonormal set in the sense that
dxχn(x,z)ϕm(x,z)=δnm,x=T,ω.

We note that as the WD may take on negative values, the eigenvalues in the expansion (13) can be negative as well. The latter circumstance distinguishes Eq. (13) from the conventional coherent-mode decomposition of the cross-correlation function [26]. Next, one can arrange the squares of the eigenvalues in decreasing order, λ02λ12λ22. Introducing the reduced eigenvalues νn(z)’s such that νn2=λn2/λ021, we can define the degree of phase-space separability of the pulse by the expression

ρ(z)=1n=0νn2(z).
It follows from the definition that the degree of separability is bound by unity, 0 ≤ ρ(z) ≤ 1, attaining its maximum if there is only the lowest-order eigenvalue present in the expansion (13). This corresponds to the ideal case of complete separability of the WD. Thus the proposed measure conforms to our intuitive perception of the degree of separability.

3. Degree of separability of chirped Gaussian Schell-model (CGSM) pulses

We will now illustrate the introduced concept using a particular example of chirped Gaussian Schell-model pulses. Not only does the latter serve as a rather representative case, but it enables us to obtain closed-form analytical results. Moreover, we can show explicitly in the case of CGSM pulses the way to control the WD’s degree of separability. CGSM pulses can be generated, for example, by transmitting GSM pulses through a time lens which imposes a quadratic phase chirp on the pulse [21].

The cross-correlation function of CGSM pulses can be written in the form

ΓCGSM(t1,t2)=Γ00exp[(1iC)t122tp2(1+iC)t222tp2(t1t2)22tc2],
where tp is a characteristic pulse width, tc is a pulse coherence time, C is a dimensionless chirp parameter, and Γ00 is a normalization constant. Using the definition of WD (3), we can express the WD of a CGSM pulse as
𝒲CGSM(t,ω)=𝒲00exp[t2tp2(1+C2teff22tp2)ω2teff22+Cteff2tp2ωt],
where
1teff2=1tp2+12tc2.
It can be readily inferred from Eq. (19) that because of phase chirping, the WD of a CGSM source is not separable.

The WD of a pulse generated by a CGSM source can be determined from Eqs. (7) and (19); the result can be represented as

𝒲CGSM(T,ω;z)=𝒲00exp[T2tp2(1+C2teff22tp2)ω2σ2(z)teff22+C(z)teff2tp2ωT].
Here we introduced the propagation factor σ(z) given by
σ2(z)=(1+Cβ2ztp2)2+2β22z2tp2teff2,
and the effective chirp parameter C(z) as
C(z)=C+β2ztp2(C2+2tp2teff2).

Eqs. (22) and (23) are generalizations to the case of partially coherent pulses of the corresponding expressions for fully coherent chirped Gaussian pulses [18]. The latter can be recovered from the former by letting teff=2tp which corresponds to tc → ∞. The evolution scenario of σ depends on the sign of 2. If 2 ≥ 0, σ increases monotonously with the propagation distance. On the other hand, if 2 < 0, the propagation factor attains a minimum,

σmin=2tp2/teff2C2+2tp2/teff2,
at the distance
z*=Cteff2/β2C2+2tp2/teff2.
This behavior is qualitatively sketched in Fig. 1. We also note that at z* the effective chirp is equal to zero–the accrued chirp on propagation in the medium unchirps the initial chirp of the opposite sign imposed by the time lens.

 figure: Fig. 1

Fig. 1 Sketching the behavior of the propagation factor of a CGSM pulse as a function of dimensionless propagation distance Z = β2z/tp.

Download Full Size | PDF

Next, we can conclude by comparing Eqs. (19) and (21) that the WD maintains its Gaussian shape, with the spectral part and the coupling term–the last term in the exponential function in Eq. (21)–scaling on propagation. Alternatively, we can work out the cross-correlation function by taking a Fourier transform of the WD with respect to ω. We arrive, after straightforward algebra, at the expression

ΓCGSM(s1,s2;z)=Γ00σ(z)exp[(s12+s22)2tp2σ2(z)(s1s2)22tc2σ2(z)+iC(z)2tp2σ2(z)(s12s22)].
It is clear from Eq. (26) that the propagation factor σ governs the dynamics of the pulse width and coherence time.

To determine the degree of separability, we use the following Mehler’s summation formula for Hermite polynomials [27]

exp(x2+y22xyζ1ζ2)=1+ζ2ex2y2n=0ζn2nn!Hn(x)Hn(y),
where |ζ| ≤ 1. Next, we introduce the scaling factors a(z) and b(z) such that in the scaled variables = t/a(z) and ω̃ = ω/b(z), the WD of a CGSM takes the form
𝒲CGSM(T˜,ω˜;z)=𝒲00exp(T˜2+ω˜22T˜ω˜ζ(z)1ζ2(z)).
On comparing Eqs. (28) and (27), we conclude that the WD in the original variables can be expanded into a bi-orthogonal series (13) with the eigenvalues
λn(z)=𝒲00[1+ζ2(z)]a(z)b(z)ζ2n(z),
and the eigenfunctions
χn(T,z)=12nn!πa(z)eT2/a2(z)Hn[Ta(z)],
and
ϕn(ω,z)=12nn!πb(z)eω2/b2(z)Hn[ωb(z)].
In Eqs. (30) and (31), the scaling factors are given by the expressions
a(z)=2tp4/teff2(C2+2tp2/teff2)[1ζ2(z)],
and
b(z)=2σ2(z)[1ζ2(z)]teff2,
where
ζ2(z)=C2(z)σ2(z)(C2+2tp2/teff2),
and positive roots are assumed in Eqs. (32) and (33). We note in passing that the bi-orthogonal decomposition (13) should not be confused with the more familiar coherent-mode expansion of a GSM source [28,29]. In our case, the eigenfunction sets {χn} and {ϕn} belong to different domains which is formally reflected in quite different scaling of a and b with the propagation distance.

Using Eqs. (34) and (29) in Eq. (17), we can show that the degree of separability turns out to be given by a remarkably simple expression

ρ(z)=σmin2/σ2(z).
The analysis of Eq. (35) reveals that the behavior of ρ as a function of the propagation distance qualitatively depends on the sign of the initial chirp. If C ≥ 0, the degree of phase separability monotonously decreases with the propagation distance. If, on the other hand, C < 0, ρ goes through a maximum attained at z = z*. These scenarios are exhibited in Fig. 2 where we present the behavior of ρ as a function of dimensionless propagation distance in the coherent case using three values of the chirp, C = 0 and C = ±1 for illustration. The influence of pulse coherence time on the evolution of ρ is illustrated in Fig. 3 for several values of tc and C = −1. Interestingly, the WD in the CGSM case becomes separable precisely at the distance where the pulses are the most compressed and least coherent. Chirping GSM pulses with a time lens then provides a powerful tool to control pulse coherence and phase-space separability at the same time.

 figure: Fig. 2

Fig. 2 Degree of phase-space separability of a fully coherent CGSM pulse as a function of dimensionless propagation distance Z = β2z/tp for three values of the initial chirp: C = 0 and C = ±1.

Download Full Size | PDF

 figure: Fig. 3

Fig. 3 Degree of phase-space separability of a partially coherent CGSM pulse as a function of dimensionless propagation distance Z = β2z/tp; solid, tc = ∞, dotted, tc = tp and dashed tc=2tp/3. The initial chirp is C = −1.

Download Full Size | PDF

We stress in conclusion that although quantitative details of the presented decomposition are specific of the CGSM pulses, qualitatively the results are quite general and entirely independent of a particular source model. Our findings can be summarized by saying that (i) the introduced degree of phase-space separability of statistical pulses is intimately related with their propagation characteristics in linear dispersive media and (ii) initial chirping, e. g., with a time lens, makes it possible to effectively control the degree of phase-space separability of the pulses.

References and links

1. T. Brabec and F. Krausz, “Intense few-cycle laser fields: Frontiers of nonlinear optics,” Rev. Mod. Phys. 72, 545–591 (2000). [CrossRef]  

2. P. Pääkkönen, J. Turunen, P. Vahimaa, A. T. Friberg, and F. Wyrowski, “Partially coherent Gaussian pulses,” Opt. Commun. 204, 53–58 (2002). [CrossRef]  

3. H. Lajunen, J. Tervo, J. Turunen, P. Vahimaa, and F. Wyrowski, “Spectral coherence properties of temporarily modulated stationary light sources,” Opt. Express 11, 1894–1899 (2003). [CrossRef]   [PubMed]  

4. S. A. Ponomarenko, G. P. Agrawal, and E. Wolf, “Energy spectrum of a nonstationary ensemble of pulses,” Opt. Lett. 29, 394–396 (2004). [CrossRef]   [PubMed]  

5. V. Torres-Company, H. Lajunen, and A. T. Friberg, “Cohrence theory of noise in ultrashort pulse trains,” J. Opt. Soc. Am. B 24, 1441–1450 (2007). [CrossRef]  

6. B. Davis, “Measurable coherence theory for statistically periodic fields,” Phys. Rev. A 76, 043843 (2007). [CrossRef]  

7. P. Vahimaa and J. Turunen, “Independent-elementary-pulse representation for non-stationary fields,” Opt. Express 14, 5007–5012 (2006). [CrossRef]   [PubMed]  

8. A. T. Friberg, H. Lajunen, and V. Torres-Company, “Spectral elementary-coherence-function representation for partially coherent light pulses,” Opt. Express 15, 5160–5165 (2007). [CrossRef]   [PubMed]  

9. J. Ojeda-Castaneda, J. Lancis, C. M. Gomez-Sarabia, V. Torres-Company, and P. Andrés, “Ambuguity function analysis of pulse train propagation: applications to temporal Lau filtering,” J. Opt. Soc. Am. A 242268–2273 (2007). [CrossRef]  

10. M. A. Alonso, “Wigner functions in optics: describing beams as ray bundles and pulses as particle ensembles,” Adv. Opt. Photon. 3, 272–365 (2011). [CrossRef]  

11. S. A. Ponomarenko, “Complex Gaussian representation of statistical pulses,” Opt. Express 19, 17086–17091 (2011). [CrossRef]   [PubMed]  

12. G. P. Agrawal, Fiber-Optic Communication Systems, 3rd ed., (Wiley, New York, NY2002). [CrossRef]  

13. Q. Lin, L. Wang, and S. Zhu, “Partially coherent light pulse and its propagation,” Opt. Commun. 219, 65–70 (2003). [CrossRef]  

14. M. Brunel and S. Coëtlemec, “Fractional-order Fourier formulation of the propagation of partially coherent light pulses,” Opt. Commun. 230, 1–5 (2004). [CrossRef]  

15. H. Lajunen, V. Torres-Company, J. Lancis, E. Silvestre, and P. Andrés, “Pulse-by-pulse method to characterize partially coherent pulse propagation in instanteneous nonlonear media,” Opt. Express 18, 14979–14991 (2011). [CrossRef]  

16. B. H. Kolner and M. Nazarathy, “Temporal imaging with a time lens,” Opt. Lett. 14, 630–632 (1989). [CrossRef]   [PubMed]  

17. Although the slowly-varying envelope approximation breaks down for few-cycle long femtosecond pulses, the decomposition into the envelope and carrier wave makes sense even in this case, see Ref. [21] for details.

18. G. P. Agrawal, Nonlinear Fiber Optics, 4th ed. (Academic Press, Amsterdam, 2007).

19. M. J. Bastiaans, “The Wigner distribution function applied to optical signals and systems,” Opt. Commun. 25, 26–30, (1978). [CrossRef]  

20. M. J. Bastiaans, “The Wigner function and its application to first-order optics,” J. Opt. Soc. Am. 69, 1710–1716 (1979). [CrossRef]  

21. J. C. Diels and W. Rudolph, Ultrashort Laser Pulse Phenomena, 2nd ed. (Academic Press, Amsterdam2006).

22. J. Lancis, V. Torres-Company, E. Silvestre, and P. Andrés, “Space-time analogy for partially coherent plane-wave-type pulses,” Opt. Lett. 30, 2973–2975 (2005). [CrossRef]   [PubMed]  

23. V. Torres-Company, J. Lancis, and P. Andrés, “Space-time analogies in optics,” Prog. Opt. 56, 1–80 (2011), ed. E. Wolf. [CrossRef]  

24. M. G. Raymer, M. Beck, and D. F. McAlister, “Complex wave-field reconstruction using phase-space tomography,” Phys. Rev. Lett. 72, 1137–1140 (1994). [CrossRef]   [PubMed]  

25. P. M. Morse and H. Feshbach, Methods of Theoretical Physics (McGraw-Hill, New York, 1953), Part I.

26. L. Mandel and E. Wolf, Optical Coherence and Quantum Optics (Cambridge University Press, 1995).

27. M. Abramowitz and I. A. Stegan, Handbook of Mathematical Functions (Dover, New York, 1972).

28. F. Gori, “Collett-Wolf sources and multimode lasers,” Opt. Commun. 34, 301–305 (1980). [CrossRef]  

29. A. Starikov and E. Wolf, “Coherent-mode representation of Gaussian Schell-model sources and their radiation fields,” J. Opt. Soc. Am. 72, 923–928 (1982). [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 (3)

Fig. 1
Fig. 1 Sketching the behavior of the propagation factor of a CGSM pulse as a function of dimensionless propagation distance Z = β2z/tp.
Fig. 2
Fig. 2 Degree of phase-space separability of a fully coherent CGSM pulse as a function of dimensionless propagation distance Z = β2z/tp for three values of the initial chirp: C = 0 and C = ±1.
Fig. 3
Fig. 3 Degree of phase-space separability of a partially coherent CGSM pulse as a function of dimensionless propagation distance Z = β2z/tp; solid, tc = ∞, dotted, tc = tp and dashed t c = 2 t p / 3. The initial chirp is C = −1.

Equations (35)

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

E ( t ) = U ( t ) e i ω c t ,
Γ ( t 1 , t 2 ) = U * ( t 1 ) U ( t 2 ) ,
𝒲 ( t , ω ) = d τ Γ ( t τ / 2 , t + τ / 2 ) e i ω τ ,
I ( t ) = d ω 𝒲 ( t , ω ) ; S ( ω ) = d t 𝒲 ( t , ω ) .
i z U 1 2 β 2 s s 2 U = 0 ,
i z Γ + 1 2 β 2 ( s 1 s 1 2 s 2 s 2 2 ) Γ = 0 .
z 𝒲 + β 2 T 𝒲 = 0 ,
T = s 1 + s 2 2 , and τ = s 1 s 2 ,
𝒲 ( T , ω ; z ) = 𝒲 0 ( T β 2 ω , ω ) .
𝒲 ( t , ω ) I ( t ) S ( ω ) .
Γ GSM ( t 1 , t 2 ) = I ( t 1 + t 2 2 ) g ( t 1 t 2 ) ,
Γ qs ( t 1 , t 2 ) I ( t 1 + t 2 2 ) γ ( t 1 t 2 ) ,
𝒲 ( T , ω ; z ) = n λ n ( z ) χ n ( T , z ) ϕ n ( ω , z ) ,
λ n ( z ) ϕ n ( ω , z ) = d T 𝒲 ( T , ω ; z ) ϕ n ( T , z ) ,
λ n ( z ) χ n ( T , z ) = d ω 𝒲 ( T , ω ; z ) χ n ( ω , z ) .
d x χ n ( x , z ) ϕ m ( x , z ) = δ n m , x = T , ω .
ρ ( z ) = 1 n = 0 ν n 2 ( z ) .
Γ CGSM ( t 1 , t 2 ) = Γ 00 exp [ ( 1 i C ) t 1 2 2 t p 2 ( 1 + i C ) t 2 2 2 t p 2 ( t 1 t 2 ) 2 2 t c 2 ] ,
𝒲 CGSM ( t , ω ) = 𝒲 00 exp [ t 2 t p 2 ( 1 + C 2 t eff 2 2 t p 2 ) ω 2 t eff 2 2 + C t eff 2 t p 2 ω t ] ,
1 t eff 2 = 1 t p 2 + 1 2 t c 2 .
𝒲 CGSM ( T , ω ; z ) = 𝒲 00 exp [ T 2 t p 2 ( 1 + C 2 t eff 2 2 t p 2 ) ω 2 σ 2 ( z ) t eff 2 2 + C ( z ) t eff 2 t p 2 ω T ] .
σ 2 ( z ) = ( 1 + C β 2 z t p 2 ) 2 + 2 β 2 2 z 2 t p 2 t eff 2 ,
C ( z ) = C + β 2 z t p 2 ( C 2 + 2 t p 2 t eff 2 ) .
σ min = 2 t p 2 / t eff 2 C 2 + 2 t p 2 / t eff 2 ,
z * = C t eff 2 / β 2 C 2 + 2 t p 2 / t eff 2 .
Γ CGSM ( s 1 , s 2 ; z ) = Γ 00 σ ( z ) exp [ ( s 1 2 + s 2 2 ) 2 t p 2 σ 2 ( z ) ( s 1 s 2 ) 2 2 t c 2 σ 2 ( z ) + i C ( z ) 2 t p 2 σ 2 ( z ) ( s 1 2 s 2 2 ) ] .
exp ( x 2 + y 2 2 x y ζ 1 ζ 2 ) = 1 + ζ 2 e x 2 y 2 n = 0 ζ n 2 n n ! H n ( x ) H n ( y ) ,
𝒲 CGSM ( T ˜ , ω ˜ ; z ) = 𝒲 00 exp ( T ˜ 2 + ω ˜ 2 2 T ˜ ω ˜ ζ ( z ) 1 ζ 2 ( z ) ) .
λ n ( z ) = 𝒲 00 [ 1 + ζ 2 ( z ) ] a ( z ) b ( z ) ζ 2 n ( z ) ,
χ n ( T , z ) = 1 2 n n ! π a ( z ) e T 2 / a 2 ( z ) H n [ T a ( z ) ] ,
ϕ n ( ω , z ) = 1 2 n n ! π b ( z ) e ω 2 / b 2 ( z ) H n [ ω b ( z ) ] .
a ( z ) = 2 t p 4 / t eff 2 ( C 2 + 2 t p 2 / t eff 2 ) [ 1 ζ 2 ( z ) ] ,
b ( z ) = 2 σ 2 ( z ) [ 1 ζ 2 ( z ) ] t eff 2 ,
ζ 2 ( z ) = C 2 ( z ) σ 2 ( z ) ( C 2 + 2 t p 2 / t eff 2 ) ,
ρ ( z ) = σ min 2 / σ 2 ( z ) .
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.