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

Modeling the depolarization of space-borne lidar signals

Open Access Open Access

Abstract

A physical model was extended with a polarization function to create a vectorized physical model (VPM) to analyze the vertical profile of the observed depolarization ratio due to multiple scattering from water clouds by space-borne lidar. The depolarization ratios due to single scattering, on-beam multiple scattering, and pulse stretching mechanisms are treated separately in the VPM. The VPM also includes a high-order scattering matrix and accommodates mechanisms that modify the polarization state during multiple scattering processes. The estimated profile of the depolarization ratio from the VPM showed good agreement with Monte Carlo simulations, with a mean relative error of about 2% ± 3%.

© 2019 Optical Society of America under the terms of the OSA Open Access Publishing Agreement

1. Introduction and physical model

The linear depolarization ratio δ from cirrus clouds mainly reflects the single scattering properties of non-spherical ice particles [1,2], due in part to the small cloud optical thickness τ at the lidar wavelength. If the particle shape and orientation can be specified, δ can be calculated by the recently developed physical optics method [3–6]. For water clouds, the cloud τ is large, and δ is affected by multiple scattering events. The δ observed by cloud-aerosol lidar with orthogonal polarization (CALIOP) [7] at 532 nm can exceed 40% due to its large footprint [8], which is larger than δ typically measured by ground-based lidars with small fields of view (FOVs). Besides the FOV, vertical cloud microphysical profiles have strong influences on the observed δ profiles [1]. A quantitative analysis of δ, in addition to the backscattering coefficient, is expected to lead to better characterization of water cloud and super-cooled microphysics, which currently possesses one of the largest uncertainties in future climate predictions [9].

To investigate the information content of δ, a double-scattering depolarization model [1,10], a triple backscattering model obtained via a time-dependent transfer model [11], and a depolarization parameter model [12] have been developed for application to ground-based lidar and multiple FOV lidar measurements. However, fewer space-borne lidar studies have analyzed δ from water clouds with regard to higher-order scattering. Based on Monte Carlo simulations for vertically homogeneous cloud profiles, ref. [13] investigated the relationship between the layer-integrated backscattering coefficient and the layer-integrated linear depolarization ratio δlayer.

The recently developed physical model (PM) [14] uses high-order scattering phase functions pn [15,16] and path integral formulation [17,18] to capture the transition of the time-dependent backscattered irradiance from the single-scattering regime to the multiple-scattering and diffusion regimes, without tracking complete scattering processes. Error analyses indicated a mean relative error of about 15% for the estimated backscattering coefficient by the PM overall, in reference to the Monte Carlo approach of [19] (referred to as MC model). Here, we introduce a vectorized physical model (VPM). Specifically, the PM [14] was extended with a polarization function to create a VPM to analyze the vertical profile of the observed depolarization ratio due to multiple scattering from water clouds by space-borne lidar.

The paper is organized as follows. In Section 2.1, the basis of the PM [14] is provided. In Section 2.2, the n-th order scattering matrix pn and the mechanisms by which the polarization state is changed according to the number of scattering events, microphysical properties, and FOV are introduced to vectorize the PM. The backscattered irradiance observed parallel and perpendicular to the incident electric field of the transmitted polarized light are estimated for single scattering events and on-beam and off-beam multiple scattering components to derive δ. In Section 3, δ is estimated using the proposed VPM and compared with results from the literature and those obtained using MC simulations. Section 4 provides a summary of our work and concluding remarks.

2. Vectorized physical model

2.1 Overview of the scalar PM

Here, we briefly describe the PM approach. In an xyz Cartesian coordinate system, the cloud top, cloud bottom, and the satellite are located at z = z1/2 ( = 0), z = zimax+1/2, and z = −zsat, respectively. The lidar onboard the satellite with beam divergence (θbd) and FOV (θfov) emits an x-polarized laser beam in the positive z direction. The cloud is divided by the scattering mean free path 1/σsca,i, where σsca,i is the scattering coefficient at cloud layer i (i = 1,2,…,imax). zi corresponds to the midpoint of the i-th cloud layer (zi-1/2 < zzi+1/2). The geometric thickness of layer Δzi equals li = 1/σsca,i, corresponding to a unit optical thickness by definition, except at the boundaries of inhomogeneity, where the optical thickness of such a layer τi ≤1 because Δzili. The total backscattered irradiance Itot is estimated at discrete time steps of 2tj (j = 1,2,…, jmax), and tj = zi = j/c, where c is the speed of light in vacuum. Subscript i = j denotes the cloud layer property corresponding to the layer indexed by an integer equal to j.

The PM decomposes Itot(tj) into three components: the single scattering component Iss(t j), the on-beam multiple scattering component Ims,tot,on(tj), and the off-beam multiple scattering component Ims,off(zi, tj), corresponding to the pulse stretching effect with z and t dependence. Thus, Itot(tj) is given by

Itot(tj)=Iss(tj)+Ims,tot,on(tj)+i=1jIms,off(zi,tj).
Iss is obtained by the single scattering approximation for a layer of finite thickness as [14]
Iss(tj)=CsIoσsca,i=jpss,i=j(π)/(4π)exp[2τ(zi=j1/2)][exp{2σext(zi=j)Δzi=j}12σext(zi=j)Δzi=j],
where σext(zi) and pss,i(θ) are the single-scattering extinction coefficient and single-scattering phase function at scattering angle θ, respectively. Cs is a factor that depends on the lidar specifications and range from the source to zi = j [14]. Multiple scattering components Ims,tot,on and Ims,off are treated similarly to Eq. (2), but with the introduction of effective extinctions σext,on and σext,off, respectively [14]:
Ims,tot,on(tj)=Jon(tj)σsca,i=jpss,i=j(π)/(4π)exp[2i=1j1(σext,on(zi)Δzi)][exp{2σext,on(zi=j)Δzi=j}12σext,on(zi=j)Δzi=j],
Ims,off(zi,tj)=k=1iIms,off,k(zi,tj)=k=1i[CsJoff(zk,zi,tj)σsca,ipn=i(π)/(4π)exp(2q=kiσext,off(zq,zk,zi,tj)Δzq)].
Jon is the source term for Ims,tot,on, which results from the ss component scattered at layer 1, as described in Fig. 1. Joff is the source term for Ims,off, resulting from the ms,on component scattered at layer k (1 ≤ ki) and incident on zi at tj. The exact expressions for Jon and Joff, as well as σext,on and σext,off, are provided in [14], and all are functions of the n-th order scattering phase function pn(θ) (n = 1, 2,…,N) [15,16]. The scattering angle θ of pn(θ) is defined with respect to the initial incident direction. Therefore, the incident direction of light after the (n−1)-th scattering event is taken to be in the z direction. Δzq´ in Eq. (4) is the photon path length in layer q, where Δzq´>Δzq due to multiple scattering.

 figure: Fig. 1

Fig. 1 (a) Relationships among ① the single scattering component Iss (red), ② the on-beam multiple scattering component Ims,tot,on (blue), and ③ the off-beam multiple scattering component Ims,off (green). The source of the ms,on component (Jon) is the ss component scattered out of the initial incident direction of the beam at layer z1. The sources of the ms,off components (Joff) are basically the ms,on components scattered at previous layers at earlier times. (b) Schematic of layer i, where zi denotes the z axis value corresponding to the center of the layer. zi-1/2 and zi + 1/2 are the bottom and top boundaries of the layer, respectively.

Download Full Size | PDF

2.2 The VPM

The basic concept of the VPM is as follows. A general expression for lidar multiple scattering of N-th order is described in [11] as an N-th order volume integration (dvN) of the product of the N set of 4 × 4 scattering phase matrices and the transformation matrix for the Stokes parameter for an angle between the (n − 1)-th and n-th scattering planes (Ln) [11]:

Itot(tj=N)=Iofv1vNLNp(θN)L3p(θ2)L2p(θ1)L1dv1dvN.
f is a function of the scattering geometry and scattering coefficients at the scattering volumes [11]. Unlike the PM, the resulting matrix in Eq. (5) is not directly related to the N-th order 4 × 4 scattering phase matrix pN. In the VPM, multiple scattering processes are modeled by considering one set of transformation matrices between the incident and the N-th scattering plane, in a manner similar to that used for a single scattering framework. In addition to introducing pN, the VPM accommodates the following two mechanisms that produce δ; (1) depolarization of the incoming wave for the N-th scattering event due to rotation of the scattering plane during multiple scattering processes and (2) geometrical effect which collects photons scattered in near backward directions at the N-th scattering event. To treat such mechanisms, Itot(tj = N) is given by:
Itot(tj=N)=IoF(LNpNL1).
The function F in Eq. (6) includes integration over the backscattering angles and represents the estimation of the geometric effects due to the ms,on and ms,off processes. It also considers the horizontal spread of the photons for the on-beam and off-beam components to determine the depolarization of the incoming wave at the layer of interest as a function of scattering order (subsection 2.2.4). Equation (6) corresponds to an extension of the formulation of the PM in Eqs. (1)-(4).

2.2.1 Introduction of a high-order Mueller matrix

It was noted in [15] that the average value of a polynomial after n scattering events equals the n-th order of the average value of the polynomials after a single scattering event when scattering is cylindrically symmetric. Based on this property, the backscattered irradiance was formulated by using the n-th order scattering phase function pn rather than tracking the complete scattering processes in [14]. pn is expressed in the Legendre polynomial expansion form with the l-th order expansion coefficient (2l + 1)(Al)n, where Al is the l-th order expansion coefficient of the single scattering phase function [14]. Hereafter we denote Al as Al,11. In this study, we consider polarimetric property of the lidar backscattering by replacing pn with the n-th order 4 × 4 scattering phase matrix pn. pn is given as follows:

pn=l=0(2l+1)Bl,nPl(cosθ),
Pl is the Legendre polynomial of the l-th order. The ij element of Bl,n is (Al,ij)n and Al,ij corresponds to the expansion coefficient of the single scattering phase matrix. In analogy to single scattering, the relationship between the scattering phase matrix pn elements and the Mueller scattering matrix Sn elements is given by Sn = σscak2pn/(4π) where k = 2π/λ and λ is the wavelength. The pn elements that are essential to Itot are shown in Fig. 2. A transition from an initial anisotropic shape to an isotropic shape occurs as the scattering order increases.

 figure: Fig. 2

Fig. 2 n-th order scattering phase matrix elements at 532 nm for a liquid particle with an effective radius (reff) of 10 µm.

Download Full Size | PDF

2.2.2 Formulation of the VPM

Here Itot(tj) is given as a function of the high-order intensity functions iw,h,js(θ), which depend on Sn as well as the polarization state of the incoming wave at the layer of interest and provide the ratio of the scattered to incident light at scattering angle θ (where subscript w stands for one of ss, ms,on, or ms,off, and subscript h = x or y).

The parallel and perpendicular components of Itot(tj) in Eq. (6) in respect to the incident electric field of the emitted light are denoted as Itot,x, and Itot,y, respectively. From Eq. (1), these components can be further expressed by Iss,h(tj), Ims,tot,on,h(tj), and Ims,off,h(zi, tj) as:

Itot(tj)=Itot,x(tj)ex+Itot,y(tj)ey=[Iss,x(tj)+Ims,tot,on,x(tj)+i=1jIms,off,x(zi,tj)]ex+[Iss,y(tj)+Ims,tot,on,y(tj)+i=1jIms,off,y(zi,tj)]ey=[Iss,x(tj)+Ims,tot,on,x(tj)+i=1jk=1iIms,off,k,x(zi,tj)]ex+[Iss,y(tj)+Ims,tot,on,y(tj)+i=1jk=1iIms,off,k,y(zi,tj)]ey.
In the VPM, Iss,h(tj), Ims,tot,on,h(tj), and Ims,off,h(zi, tj) in Eq. (8) are expressed in terms of iw,h,js as follows:
Iss,h(tj)=CsIo1k2iss,hs(π)exp[2τ(zi=j1/2)][exp{2σext(zi=j)Δzi=j}12σext(zi=j)Δzi=j](h=xory),
Ims,tot,on,h(tj)=CsJon(tj)Con1k2θmin,on,jθmax,on,jion,h,js(θbk)pss,i=j(θin)sinθindθinθmin,on,jθmax,on,jpss,i=j(θin)sinθindθinexp[2i=1j1(σext,on(zi)Δzi)][exp{2σext,on(zi=j)Δzi=j}12σext,on(zi=j)Δzi=j],
Ims,off,h(zi,tj)=k=1iIms,off,k,h(zi,tj)=k=1i[CsJoff(zk,zi,tj)Coff,k,i,j1k2θmin,,off,jθmax,off,jioff,h,js(θbk)pn=j(θin)sinθindθinθmin,,off,jθmax,off,jpn=j(θin)sinθindθinexp(2q=kiσext,off(zq,zk,zi,tj)Δzq)].
iw,h,js are given as a function of Sn and the depolarization of the incoming wave in subsection 2.2.3. The decay rate is assumed to be independent of the polarization state. Con and Coff,k,i,j are proportionality factors to ensure the same Itot for the PM and VPM. The angular integrations over θin are introduced to accommodate the geometric effect. The backscattering angle to the detector θbk in Eqs. (10) and (11) is given by the incident angle θin and the geometric position of the cloud layer and satellite [10], as shown in Fig. 3 as
θbk=π(θin(tj)θsat),
θsat=tan1[(zizi1)tan(θin(tj))/(zsat+zi)].
θin will be discussed in subsections 2.2.4.1 and 2.2.4.2.

 figure: Fig. 3

Fig. 3 Geometry for the definition of θbk in Eq. (12).

Download Full Size | PDF

2.2.3 Procedure to estimate the parallel and perpendicular components of irradiance

In the following, we consider the electric fields used to estimate Iss,h(tj), Ims,tot,on,h(tj), and Ims,off,h(zi, tj).

iw,h,jsin Eqs. (9)–(11) are the x and y components of the azimuthally integrated normalized intensity functions between θ and θ + dθ (dθ → 0) of a concentric cone [20], as shown in Fig. 4, defined as:

iw,h,js(θ)=|Ew,h,js|2ϕ/(|Ew,x,jin|2+|Ew,y,jin|2).
< >ϕ indicates average over 0<ϕ≤2π. Ew,js represents the scattered electric fields at the n = j-th scattering event and is related to the incident fields after the n = (j − 1)-th scattering event Ew,jin by the following [10,20]:
Ew,js=LϕDsn=jLϕEw,jin.
Here, Lϕ=(cosϕsinϕsinϕcosϕ) is used to work in the scattering plane, D=(cosθ001) expresses the fact that the perpendicular polarization is maintained in the process, while the paralell component is obtained by projection of the scattered light vector in the xy plane, and sn is the n-th order amplitude scattering matrix. The multiplicative factor [21] is omitted in Eq. (15).

 figure: Fig. 4

Fig. 4 General geometries of the incident and scattering planes. θ is the zenith angle from the incident direction z and ϕ is the azimuthal angle. The parallel and perpendicular components of the scattered electric fields are taken in the xy plane.

Download Full Size | PDF

In this study, we considered randomly oriented particles in space. Therefore, the Mueller matrix Sn can be written as [21]:

Sn=[S11,n(θ)S12,n(θ)00S12,n(θ)S22,n(θ)0000S33,n(θ)S34,n(θ)00S34,n(θ)S44,n(θ)]
From Eqs. (14)–(16), the intensity functions are derived as follows [20]:
iw,x,js(θ)=(m11,n+m22,n)|Ew,x,jin|2+(m11,nm22,n)|Ew,y,jin|2|Ew,jin|2=(m11,n+m22,n)+(m11,nm22,n)Qw, jin1+Qw, jin(n=j),
iw,y,js(θ)=(m11,nm22,n)|Ew,x,jin|2+(m11,n+m22,n)|Ew,y,jin|2|Ew,jin|2=(m11,nm22,n)+(m11,n+m22,n)Qw, jin1+Qw, jin(n=j),
where
Qw, jin=|Ew, y,jin|2/|Ew, x, jin|2,(w=ss,on,off)
are the depolarization of the incoming wave at the layer of interest, and Qss, jin=0. The parallel and perpendicular components in Eqs. (17) and (18) are given by:
m11,n+m22,n=A(θ)[2S11,n(θ)+S22,n(θ)]+3B(θ)S12,n(θ)2C(θ)S33,n(θ),
m11,nm22,n=A(θ)[2S11,n(θ)S22,n(θ)]+B(θ)S12,n(θ)+2C(θ)S33,n(θ).
The coefficients A, B, and C are as follows:
A(θ)=(1+cos2θ)/8,
B(θ)=sin2θ/8,
C(θ)=cosθ/8.
For x-polarized light, when single scattering by spherical particles (s3 = s4 = 0) is considered, Eqs. (17) and (18) reduce to the familiar expressions given in Eqs. (A15) and (A14) of [10], respectively. Equations (17) and (18) describe the relationships among Iss,h(tj), Ims,tot,on,h(tj), and Ims,off,h(zi, tj) in Eqs. (9)–(11) and Sn. Note that, for ss and ms,on, S for single scattering is used for m11 and m22 at a backscattering angle of θbk in Eqs. (17) and (18), to estimate iw,h,js(θbk) in Eqs. (9) and (10).

Once we obtain Qw, jin in Eqs. (17) and (18) as well as θin in Eq. (12), Iss,h(tj), Ims,tot,on,h(tj), and Ims,off,h(zi, tj) can be derived.

2.2.4 Depolarization of the incoming wave

In the VPM, mechanisms to produce non zero Qw, jin due to rotation of the scattering plane without tracking every scattering event as in Monte Carlo method are introduced separately for ms,on and ms,off. Here we describe how Qon,jin and Qoff,jin are produced and evlove with scattering order.

2.2.4.1 On-beam scattering

Qon,jinis estimated by considering a mean forward incident angle θref determined by the mean horizontal spread of the photons after n = j-th scattering event and FOV. To obtain Qon,jin, depolarization of the incoming wave at each cloud layer Δj,j(1≤ jj′-1) contributing to Qon,jin is estimated by the following recurrenece relation derived by using Eqs. (17) and (18):

Δj+1,j=[m11,n(θ)m22,n(θ)]+[m11,n(θ)+m22,n(θ)]Δj,j[m11,n(θ)+m22,n(θ)]+[m11,n(θ)m22,n(θ)]Δj,j(j1j1,n=j),
From Eq. (25), Δj,jis obtained by the recurrence relation beginning with Δ1,j=0 and with a constant θ = θref throughout the recurrence computation.
θref=tan1[rref,i/li](i=j),
where index i in Eq. (26) is fixed to the maximum j-index value of the sequence j′. li is the scattering mean free path and rref,i is taken to be the minimum of half rfov or the mean horizontal spread of the photons, as follows:

rref,i=min[0.5rfov(zi),k=0i1(lk+11gn=k2)](i=j).

rfov(zi) is the FOV footprint sizes at zi. gn is the n-th order asymmetry factor [14]. Finally,

Qon,jin=Δj,jδjj,
and δjjis the Kronecker delta. Qon,jinfor different θref and scattering order is shown in Fig. 5. For a given θref, Qon,jinincreases with scattering order, i.e., optical thickness. For constant scattering order, Qon,jin increases with θref, since larger θref indicates larger FOV or σext. At small incident scattering angles, Qon,jindoes not develop much during near-forward multiple scattering events.

 figure: Fig. 5

Fig. 5 Depolarization of the incoming wave for ms,on Qon,jinas a function of θ ref and the scattering order n ( = j) for reff = 10 µm at λ = 532 nm.

Download Full Size | PDF

To determine the geometrical effect for ms,on, the minimum and maximum ranges of θin in Eq. (10) are estimated from the FOV and by the diffraction widths θdf as follows:

θmax,on,j=tan1[rfov(zi)/li](i=j),
θmin,on,j={θdf(θmax,on,j>θdf)0(θmax,on,jθdf).

θmax,on,j is depicted in Fig. 6. In Eq. (30), the second minima of the diffraction wave is considered for θdf, where θdf = sin−1(2.22λ/2reff) [22,23], and reff is the effective radius [6].

 figure: Fig. 6

Fig. 6 Geometry of defining θmax,on,j and θref, where i = j and when ref,i in Eq. (27) equaled 0.5rfov.

Download Full Size | PDF

2.2.4.2 Off-beam scattering

Qoff,jinis estimated similarly toQon,jin, but by considering the horizonal spread of the photons scattered at large scattering angles due to mutiple scattering for the off-beam component. Qoff,jinis defined by the following integration over 0≤θ ≤ π as:

Qoff,j+1in=0π{(m11,nm22,n)+(m11,n+m22,n)Qoff,jin}pn(θ)sinθdθ0π{(m11,n+m22,n)+(m11,nm22,n)Qoff,jin}pn(θ)sinθdθ(j1j1,n=j),
where Qoff,1in=0 and j′ is the maximum j-index value of interest. Qoff,jin in Eq. (31), shown in Fig. 7, gradually increases with the scattering order.

 figure: Fig. 7

Fig. 7 Depolarization of the incoming wave for ms,off Qoff,jinas a function of scattering order n ( = j) for reff = 10 µm at λ = 532 nm.

Download Full Size | PDF

To determine the geometrical effect for ms,off, the range of θin in Eq. (11) is specified in the following. The source term Joff(zk,zi, tj) for Ims,off,h,k(zi, tj) in Eq. (11) is determined from the fraction of the ms,on component scattered at zk within the angle θoff,ki(tj-1) < θ < θoff,ki(tj) [14]. θoff,ki(tj) is defined in [14] as:

θoff,ki(tj)=cos1[(zizk)/{(tj+1/2czk)}].
The range of θin is given by:
θmin,off,j=θoff,ki(tj1),
θmax,off,j=θoff,ki(tj).
It is noted that Qon,jin estimated at θref = θin is taken as Qoff,jin when it is larger than Qoff,jingiven by Eq. (31).

In summary, to estimate Iss,h(tj), Ims,tot,on,h(tj), and Ims,off,h(zi, tj), we first estimate the ratios of the parallel and perpendicular components of the incoming wave at the n = j-th scattering event Qw,jin (w = ss, on, off) using Eq. (28) and Eq. (31) to calculate the corresponding normalized intensity functions iw,h,js(θbk) from Eqs. (17) and (18). iw,h,js(θbk)are then substituted into Eqs. (9)–(11) to obtain Iss,h(tj), Ims,tot,on,h(tj), and Ims,off,h(zi,tj). Finally, the depolarization ratio δ(tj) is obtained by substituting these components into Eq. (35):

δ(tj)=Itot,y(tj)Itot,x(tj)=Iss,y(tj)+Ims,tot,on,y(tj)+i=1jIms,off,y(zi,tj)Iss,x(tj)+Ims,tot,on,x(tj)+i=1jIms,off,x(zi,tj).

3. Simulation of δ with the VPM for spaceborne lidar

The applicability of the VPM to lidar is studied in three subsections: (1) the Lidar In-space Technology Experiment (LITE) [24] case, (2) ground-based lidar case, and (3) CALIOP case and characterization of δ in terms of microphysics and FOVs obtained by the VPM (δVPM) with respect to MC (δMC) [19].

3.1 LITE case

Figure 8(a) shows the normalized backscattered returns from the C1 cloud [25] for the LITE case (λ = 1064 nm), in which θFOV ranged from 0.2 to 3.5mrad, as described and simulated by The National Aeronautical and Space Administration’s Langley Monte Carlo model in [24]. Simulations using the MC model and VPM are also provided for the same cases. Based on Monte Carlo simulations, ref [13] derived a relationship between the layer-integrated multiple scattering factor η and the layer-integrated linear depolarization ratio δlayer. η is the ratio of the layer-integrated single-to-total attenuated backscattering coefficients; η becomes smaller as multiple scattering increases relative to the single-scattering component. The η factor is a useful concept for deriving cloud microphysics [26]. The MC model showed good agreement with the results of [24] and [13] in terms of the backscattered returns and ηδlayer relation for the C1 cloud at smaller δlayer; MC results in Fig. 8(c) showed a slightly larger value for η given the same δlayer and VPM overlaid the MC results (not shown).

 figure: Fig. 8

Fig. 8 (a) Normalized backscattered returns estimated using Monte Carlo (MC) [19] simulations (black lines) and the vectorized physical model (VPM; blue symbols), and produced by [24] (gray open squares and vertical bars). Note that the values of the gray open squares and the range of the literature values shown by vertical bars were extracted from Fig. 3 of [24]. (b) Linear depolarization ratio δ estimated by the VPM and MC simulations corresponding to (a) for the C1 cloud [25] at 1064 nm for the Lidar In-space Technology Experiment (LITE) case with zsat = 293 km. (c) Layer-integrated ηδlayer relations by [13] (black line) and by MC [19] (colored lines).

Download Full Size | PDF

In Figs. 8(a) and 8(b), MC results show larger backscattered returns and δMC values for the larger FOV cases. A comparison of δVPM and δMC showed that these trends were captured well by the VPM. The mean relative error of the VPM results was less than about 3.6% overall, with respect to that of MC simulations.

3.2 Ground-based lidar case

A simulation was also performed for a ground-based lidar case to test the applicability of the VPM. In Fig. 9, δVPM is compared with δMC and literature values of δ obtained by other models reported in [27], in which a ground-based lidar (λ = 700 nm; θFOV ~1.75mrad) located 2 km from the cloud base was considered. δ from other models [27] (i.e., the 3DMcPOLID model, the ECSIM model, and a semi-analytical approach with small-angle approximation, described in [27] and in the references therein) had about 10% relative variability in the predicted range. The δMC generally showed good agreement with other models in [27] and was located at the upper boundary of the variability range, corresponding to the results obtained with the 3DMcPOLID model [27]. The VPM and MC showed good consistency, with the exception at small τ, where δVPM was within the variability range but slightly underestimated δMC.

 figure: Fig. 9

Fig. 9 δ from the C1 cloud [25] obtained by VPM and MC for a ground-based lidar at 700 nm. Ranges of the mean δ values predicted by different methods described in [27] and as shown by vertical bars in the figure were extracted from Fig. A4 of [27].

Download Full Size | PDF

3.3 CALIOP case and VPM characterization

Sensitivity studies of δVPM were performed, following [14]. The accuracy of the VPM was investigated by conducting 17 experiments with varying parameter settings (Table 1). The relative errors (|ERR|) of the VPM with respect to MC were calculated as |ERR| = (|δVPM−δMC|)/δMC. The overall relative errors 〈ERR〉 (%) were calculated for different τ ranges by averaging |ERR| estimated at all data points.

Tables Icon

Table 1. Experimental setting.

For Experiments 1–13, the reference simulations were performed using CALIOP specifications (λ = 532 nm; θfov,1 = 0.13mrad; 702km to the cloud top) [14]. Experiments 1–12 corresponded to homogeneous cloud profiles with different σext, reff, and FOVs [σext = 3, 15.7, and 40km−1; reff = 5, 10, and 20µm; and θfov = θfov,1 and 10θfov,1]. A comparison of δVPM and δMC for the abovementioned experiments is shown in Fig. 10(b). δVPM showed good agreement with δMC, with 〈ERR〉 values of less than about 3.2% on average.

 figure: Fig. 10

Fig. 10 (a) δ obtained by VPM (δVPM) and MC (δMC) for Experiment 6. In the figure, δVPM estimated without considering the depolarization of the incoming wave (δPn) is also shown for Experiment 6 (black star). (b) δVPM and δMC for Experiments 2, 4, 6, 8, 10, and 12.

Download Full Size | PDF

To analyze the importance of considering the depolarization of the incoming wave on δ, δVPM with Qw,jin = 0 (labeled as δPn) is presented in Fig. 10(a) for Experiment 6. δPn is the δ produced by pn and the geometrical effect. δPn experiment underestimated δMC from τ = 3 (z~191m). Without considering Qw,jin, the difference δMCδPn was about 0.2 at τ = 8. In [14], Itot was categorized into ss-dominant, ms,on-dominant, and ms,off-dominant regimes, bounded at about τ = 2 (z~127m) and τ = 8 (z~510m) for Experiment 6. The slope of δPn also changed according to the three regimes; however, much better agreement between δVPM and δMC was achieved when Qw,jin was considered.

The dependence of δ on reff for a given σext is further examined in Experiments 1–3, 5–7, and 9–11 for σext = 3km−1, σext~15.7km−1, and σext = 40km−1, respectively. The δ for reff = 5, 10, and 20µm (denoted as δ5, δ10, and δ20, respectively) are shown in Figs. 11(a-c), where it is seen that the sensitivity of δ to reff decreased as τ increased for a constant σext. The difference |δ5-δ20| was about 0.12, 0.12, and 0.1 at τ = 2, τ = 5, and τ = 10 for σext = 3km−1, respectively. These differences corresponded to a relative difference Δδ = |δ5-δ20|/δ20 of about 63%, 37%, and 1.2% for τ = 2, τ = 5, and τ = 10, respectively. Both MC and VPM also indicated that the sensitivity of δ to reff decreased as σext increased for a given τ. For σext~15.7km−1, |δ5-δ20| was about 0.11, 0.11, and 0.06 corresponding to Δδ of about 39%, 23%, and 8% at τ = 2, τ = 5, and τ = 10, respectively. In the case of σext = 40km−1, |δ5-δ20| and Δδ were about 0.08, 0.09, 0.05 and 23%, 17%, 6% at τ = 2, τ = 5, τ = 10, respectively. From Figs. 11(a-c), the retrieval uncertainty in τ by δ due to the variation in reff and σext can also be inferred. It was found that when δ = 0.3, the corresponding τ had about 2 times, 1.7 times, and 1.4 times difference among reff considered at σext = 3km−1, σext~15.7km−1, and σext = 40km−1, respectively. At δ = 0.7, there was about 1.1~1.2 times difference in τ despite the σext considered.

 figure: Fig. 11

Fig. 11 δVPM and δMC for the (a) σext = 3km−1 case corresponding to Experiments 1,2 and 3, (b) σext~15.7km−1 case corresponding to Experiments 5,6 and 7, (c) σext = 40km−1 case corresponding to Experiments 9,10 and 11.

Download Full Size | PDF

Experiment 13 corresponds to the case with vertically inhomogeneous cloud particle size profiles, specifically, reff = 20, 5, and 10 µm from the cloud top bounded at 240 and 480 m with a liquid water content of 0.1g/m3. 240m corresponded to the vertical resolution of the global cloud phase product of [8] (KU product). Inhomogeneous structures were evident due to the change in microphysics, as observed in δMC in Fig. 12. Such behavior was well captured by δVPM; the 〈ERR〉 value was 3.76% ± 3.22%.

 figure: Fig. 12

Fig. 12 δVPM and δMC for the inhomogeneous case (Experiment 13). τ is around 1.8 and 9.5 at z = 240m and z = 480m, respectively.

Download Full Size | PDF

Experiments 14–16 and 17 corresponded to the cases studied in Figs. 8 and 9, respectively. The 〈ERR〉 values for all 17 cases are summarized in Fig. 13. The overall 〈ERR〉 was 2.32% on average, and was less than 5% for all optical thickness ranges.

 figure: Fig. 13

Fig. 13 Relative error of δVPM with respect to δMC for the 17 experiments summarized in Table 1. The results are categorized according to the optical thickness: (a) 0 ≤ τ, (b) τ ≤ 2, (c) 2 < τ < 8, and (d) 8 ≤ τ. The maximum τ values considered are the same as those used in [14] for Experiments 1–13; about 48 were used for the σext = 40 km−1 cases and about 18 were used for the others. The overall mean relative errors 〈ERR〉 and standard deviations (%) listed in the upper right corner of each figure panel are estimated from all 17 profiles.

Download Full Size | PDF

4. Summary

In our proposed method, the PM developed in [14] for the total backscattered irradiances Itot from spaceborne lidars is vectorized to create a VPM to simulate polarization properties under multiple scattering conditions effectively for the first time. The VPM models vectorized lidar multiple scattering with one set of n-th order Mueller matrices Sn and a transformation matrix for Stokes parameters L connecting the initial incident plane and n-th scattering plane. Additional treatment was introduced to implement the accumulated effects on the depolarization. This consists of two mechanisms: (1) depolarization of the incoming wave at the cloud layer of interest during multiple scattering in the forward direction due to rotation of the scattering plane and (2) depolarization produced by angular distributions of the backward scattering processes within the FOV (geometrical effect). The depolarization of the incoming wave was determined by considering the horizontal spread of the photons for the on-beam and off-beam components. These mechanisms were modeled separately for Iss(tj), Ims,tot,on(tj), and Ims,off(zi, tj) to estimate their parallel and perpendicular components in respect to the incident electric field of the emitted light to obtain δ(tj). These formulations accommodate the dependence of δ on the scattering order, microphysical properties, and FOV.

Evaluation of δ from VPM against MC simulations and, in part, by literature values was performed for a total of 17 cases. The effect of the depolarization of the incoming wave became important as τ increased. The comparisons showed overall agreement between the VPM and MC simulations, with a mean relative error of about 2% ± 3%. The VPM was more efficient than the MC approach. The sensitivity of δ to reff was also investigated, where the relative difference in δ among reff considered was generally larger than the relative error of the VPM against MC.

In future studies, the dependence of the vertical profiles of δ and Itot on cloud microphysics will be investigated comprehensively using the VPM, based on measurements from space-borne lidars [7,28], as well as ground-based multiple scattering lidars [29], for satellite algorithms. The extension of the VPM to circular polarization used in multiple FOV measurements [10] should be a straightforward process.

Funding

Japan Society for the Promotion of Science (JSPS KAKENHI Grant Numbers JP18K03745, JP17H06139); The Japan Aerospace Exploration Agency (EarthCARE Research Announcement); Ministry of Education, Culture, Sports, Science and Technology (The Arctic Challenge for Sustainability (ArCS)); Collaborated Research Program of Research Institute for Applied Mechanics, Kyushu University (Fukuoka, Japan).

References

1. K. Sassen and H. Zhao, “Lidar multiple scattering in water droplets clouds: toward and improved treatment,” Opt. Rev. 2(5), 394–400 (1995). [CrossRef]  

2. S. Reichardt and J. Reichardt, “Effect of multiple scattering on depolarization measurements with spaceborne lidars,” Appl. Opt. 42(18), 3620–3633 (2003). [CrossRef]   [PubMed]  

3. A. Borovoi, A. Konoshonkin, N. Kustova, and H. Okamoto, “Backscattering Mueller matrix for quasi-horizontally oriented ice plates of cirrus clouds: application to CALIPSO signals,” Opt. Express 20(27), 28222–28233 (2012). [CrossRef]   [PubMed]  

4. A. Konoshonkin, A. Borovoi, N. Kustova, H. Okamoto, H. Ishimoto, Y. Grynko, and J. Förstner, “Light scattering by ice crystals of cirrus clouds: From exact numerical methods to physical-optics approximation,” J. Quant. Spectrosc. Radiat. Transf. 195, 132–140 (2017). [CrossRef]  

5. H. Okamoto, K. Sato, and Y. Hagihara, “Global analysis of ice microphysics from CloudSat and CALIPSO: Incorporation of specular reflection in lidar signals,” J. Geophys. Res. 115(D22), D22209 (2010).

6. K. Sato and H. Okamoto, “Refinement of global ice microphysics using spaceborne active sensors,” J. Geophys. Res. 116(D20), D20202 (2011). [CrossRef]  

7. D. M. Winker, M. A. Vaughan, A. Omar, Y. Hu, K. A. Powell, Z. Liu, W. H. Hunt, and S. A. Young, “Overview of the CALIPSO Mission and CALIOP Data Processing Algorithms,” J. Atmos. Ocean. Technol. 26(11), 2310–2323 (2009). [CrossRef]  

8. R. Yoshida, H. Okamoto, Y. Hagihara, and H. Ishimoto, “Global analysis of cloud phase and ice crystal orientation from Cloud-Aerosol Lidar and Infrared Pathfinder Satellite Observation (CALIPSO) data using attenuated backscattering and depolarization ratio,” J. Geophys. Res. 115, D00H32 (2010). [CrossRef]  

9. S. Bony and J. L. Dufresne, “Marine boundary layer clouds at the heart of tropical cloud feedback uncertainties in climate models,” Geophys. Res. Lett. 32(20), L20806 (2005). [CrossRef]  

10. G. Roy and N. Roy, “Relation between circular and linear depolarization ratios under multiple-scattering conditions,” Appl. Opt. 47(35), 6563–6579 (2008). [CrossRef]   [PubMed]  

11. Q. Cai and K. N. Liou, “Theory of Time-Dependent Multiple Backscattering from Clouds,” J. Atmos. Sci. 38(7), 1452–1466 (1981). [CrossRef]  

12. G. Roy, G. Tremblay, and X. Cao, “Scattering phase function depolarization parameter model and its application to water droplets sizing using off-axis lidar measurements at multiple angles,” Appl. Opt. 57(4), 969–977 (2018). [CrossRef]   [PubMed]  

13. Y. Hu, “Depolarization ratio-effective lidar ratio relation: Theoretical basis for space lidar cloud phase discrimination,” Geophys. Res. Lett. 34(11), L11812 (2007). [CrossRef]  

14. K. Sato, H. Okamoto, and H. Ishimoto, “Physical model for multiple scattered space-borne lidar returns from clouds,” Opt. Express 26(6), A301–A319 (2018). [CrossRef]   [PubMed]  

15. S. Goudsmit and J. L. Saunderson, “Multiple Scattering of Electrons,” Phys. Rev. 57(1), 24–29 (1940). [CrossRef]  

16. H. W. Lewis, “Multiple scattering in an infinite medium,” Phys. Rev. 78(5), 526–529 (1950). [CrossRef]  

17. L. T. Perelman, J. Wu, I. Itzkan, and M. S. Feld, “Photon migration in turbid media using path integrals,” Phys. Rev. Lett. 72(9), 1341–1344 (1994). [CrossRef]   [PubMed]  

18. R. P. Feynman, “Space-Time Approach to Quantum Electrodynamics,” Phys. Rev. 76(6), 769–789 (1949). [CrossRef]  

19. H. Ishimoto and K. Masuda, “A Monte Carlo approach for the calculation of polarized light: application to an incident narrow beam,” J. Quant. Spectrosc. Radiat. Transf. 72(4), 467–483 (2002). [CrossRef]  

20. A. Ben-David, “Mueller matrices and information derived from linear polarization lidar measurements: theory,” Appl. Opt. 37(12), 2448–2463 (1998). [CrossRef]   [PubMed]  

21. C. F. Bohren and D. R. Huffman, Absorption and Scattering of Light by Small Particles (Wiley, 1983).

22. K. N. Liou, An Introduction to Atmospheric Radiation (Academic Press, 2002).

23. M. Born and E. Wolf, Principles of optics, seventh (expanded) edition (Cambridge university press, 1999).

24. D. M. Winker and L. R. Poole, “Monte-Carlo calculations of cloud returns for ground-based and space-based lidars,” Appl. Phys. B 60(4), 341–344 (1995). [CrossRef]  

25. D. Deirmendjian, Electronmagnetic Scattering on Spherical Poly-dispersions (Elsevier, 1969)

26. C. M. Platt, “Remote sounding of High Clouds. III: Monte Carlo Calculations of Multiple-Scattered Lidar Returns,” J. Atmos. Sci. 38(1), 156–167 (1981). [CrossRef]  

27. A. Alkasem, F. Szczap, C. Cornet, V. Shcherbakov, Y. Gour, O. Jourdan, L. C. Labonnote, and G. Mioche, “Effects of cirrus heterogeneity on lidar CALIOP/CALIPSO data,” J. Quant. Spectrosc. Radiat. Transf. 202, 38–49 (2017). [CrossRef]  

28. A. J. Illingworth, H. W. Barker, A. Beljaars, M. Ceccaldi, H. Chepfer, N. Clerbaux, J. Cole, J. Delanoë, C. Domenech, D. P. Donovan, S. Fukuda, M. Hirakata, R. J. Hogan, A. Huenerbein, P. Kollias, T. Kubota, T. Nakajima, T. Y. Nakajima, T. Nishizawa, Y. Ohno, H. Okamoto, R. Oki, K. Sato, M. Satoh, M. W. Shephard, A. Velázquez-Blázquez, U. Wandinger, T. Wehr, and G.-J. van Zadelhoff, “The EarthCARE Satellite: The Next Step Forward in Global Measurements of Clouds, Aerosols, Precipitation, and Radiation,” Bull. Am. Meteorol. Soc. 96(8), 1311–1332 (2015). [CrossRef]  

29. H. Okamoto, K. Sato, T. Nishizawa, N. Sugimoto, T. Makino, Y. Jin, A. Shimizu, T. Takano, and M. Fujikawa, “Development of a multiple-field-of-view multiple-scattering polarization lidar: comparison with cloud radar,” Opt. Express 24(26), 30053–30067 (2016). [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 (13)

Fig. 1
Fig. 1 (a) Relationships among ① the single scattering component Iss (red), ② the on-beam multiple scattering component Ims,tot,on (blue), and ③ the off-beam multiple scattering component Ims,off (green). The source of the ms,on component (Jon) is the ss component scattered out of the initial incident direction of the beam at layer z1. The sources of the ms,off components (Joff) are basically the ms,on components scattered at previous layers at earlier times. (b) Schematic of layer i, where zi denotes the z axis value corresponding to the center of the layer. zi-1/2 and zi + 1/2 are the bottom and top boundaries of the layer, respectively.
Fig. 2
Fig. 2 n-th order scattering phase matrix elements at 532 nm for a liquid particle with an effective radius (reff) of 10 µm.
Fig. 3
Fig. 3 Geometry for the definition of θbk in Eq. (12).
Fig. 4
Fig. 4 General geometries of the incident and scattering planes. θ is the zenith angle from the incident direction z and ϕ is the azimuthal angle. The parallel and perpendicular components of the scattered electric fields are taken in the xy plane.
Fig. 5
Fig. 5 Depolarization of the incoming wave for ms,on Q on, j in as a function of θ ref and the scattering order n ( = j) for reff = 10 µm at λ = 532 nm.
Fig. 6
Fig. 6 Geometry of defining θmax,on,j and θref, where i = j and when ref,i in Eq. (27) equaled 0.5rfov.
Fig. 7
Fig. 7 Depolarization of the incoming wave for ms,off Q off,j in as a function of scattering order n ( = j) for reff = 10 µm at λ = 532 nm.
Fig. 8
Fig. 8 (a) Normalized backscattered returns estimated using Monte Carlo (MC) [19] simulations (black lines) and the vectorized physical model (VPM; blue symbols), and produced by [24] (gray open squares and vertical bars). Note that the values of the gray open squares and the range of the literature values shown by vertical bars were extracted from Fig. 3 of [24]. (b) Linear depolarization ratio δ estimated by the VPM and MC simulations corresponding to (a) for the C1 cloud [25] at 1064 nm for the Lidar In-space Technology Experiment (LITE) case with zsat = 293 km. (c) Layer-integrated ηδlayer relations by [13] (black line) and by MC [19] (colored lines).
Fig. 9
Fig. 9 δ from the C1 cloud [25] obtained by VPM and MC for a ground-based lidar at 700 nm. Ranges of the mean δ values predicted by different methods described in [27] and as shown by vertical bars in the figure were extracted from Fig. A4 of [27].
Fig. 10
Fig. 10 (a) δ obtained by VPM (δVPM) and MC (δMC) for Experiment 6. In the figure, δVPM estimated without considering the depolarization of the incoming wave (δPn) is also shown for Experiment 6 (black star). (b) δVPM and δMC for Experiments 2, 4, 6, 8, 10, and 12.
Fig. 11
Fig. 11 δVPM and δMC for the (a) σext = 3km−1 case corresponding to Experiments 1,2 and 3, (b) σext~15.7km−1 case corresponding to Experiments 5,6 and 7, (c) σext = 40km−1 case corresponding to Experiments 9,10 and 11.
Fig. 12
Fig. 12 δVPM and δMC for the inhomogeneous case (Experiment 13). τ is around 1.8 and 9.5 at z = 240m and z = 480m, respectively.
Fig. 13
Fig. 13 Relative error of δVPM with respect to δMC for the 17 experiments summarized in Table 1. The results are categorized according to the optical thickness: (a) 0 ≤ τ, (b) τ ≤ 2, (c) 2 < τ < 8, and (d) 8 ≤ τ. The maximum τ values considered are the same as those used in [14] for Experiments 1–13; about 48 were used for the σext = 40 km−1 cases and about 18 were used for the others. The overall mean relative errors 〈ERR〉 and standard deviations (%) listed in the upper right corner of each figure panel are estimated from all 17 profiles.

Tables (1)

Tables Icon

Table 1 Experimental setting.

Equations (35)

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

I tot ( t j )= I ss ( t j )+ I ms,tot,on ( t j )+ i=1 j I ms,off ( z i , t j ) .
I ss ( t j )= C s I o σ sca,i=j p ss,i=j ( π )/ (4π) exp[ 2τ( z i=j1/2 ) ][ exp{ 2 σ ext ( z i=j )Δ z i=j }1 2 σ ext ( z i=j )Δ z i=j ],
I ms,tot,on ( t j )= J on ( t j ) σ sca,i=j p ss,i=j ( π )/ (4π) exp[ 2 i=1 j1 ( σ ext,on ( z i )Δ z i ) ][ exp{ 2 σ ext,on ( z i=j )Δ z i=j }1 2 σ ext,on ( z i=j )Δ z i=j ],
I ms,off ( z i , t j )= k=1 i I ms,off,k ( z i , t j ) = k=1 i [ C s J off ( z k , z i , t j ) σ sca,i p n=i ( π )/ (4π) exp( 2 q=k i σ ext,off ( z q , z k , z i , t j )Δ z q ) ] .
I tot ( t j=N )= I o f v 1 v N L N p( θ N ) L 3 p( θ 2 ) L 2 p( θ 1 ) L 1 d v 1 d v N .
I tot ( t j=N )= I o F( L N p N L 1 ).
p n = l=0 ( 2l+1 ) B l, n P l ( cosθ ),
I tot ( t j )= I tot,x ( t j ) e x + I tot,y ( t j ) e y =[ I ss,x ( t j )+ I ms,tot,on,x ( t j )+ i=1 j I ms,off,x ( z i , t j ) ] e x +[ I ss,y ( t j )+ I ms,tot,on,y ( t j )+ i=1 j I ms,off,y ( z i , t j ) ] e y =[ I ss,x ( t j )+ I ms,tot,on,x ( t j )+ i=1 j k=1 i I ms,off,k,x ( z i , t j ) ] e x +[ I ss,y ( t j )+ I ms,tot,on,y ( t j )+ i=1 j k=1 i I ms,off,k,y ( z i , t j ) ] e y .
I ss,h ( t j )= C s I o 1 k 2 i ss,h s ( π )exp[ 2τ( z i=j1/2 ) ][ exp{ 2 σ ext ( z i=j )Δ z i=j }1 2 σ ext ( z i=j )Δ z i=j ]( h=xory ),
I ms,tot,on,h ( t j )= C s J on ( t j ) C on 1 k 2 θ min,on,j θ max,on,j i on,h,j s ( θ bk ) p ss,i=j ( θ in )sin θ in d θ in θ min,on,j θ max,on,j p ss,i=j ( θ in )sin θ in d θ in exp[ 2 i=1 j1 ( σ ext,on ( z i )Δ z i ) ][ exp{ 2 σ ext,on ( z i=j )Δ z i=j }1 2 σ ext,on ( z i=j )Δ z i=j ],
I ms,off,h ( z i , t j )= k=1 i I ms,off,k,h ( z i , t j ) = k=1 i [ C s J off ( z k , z i , t j ) C off,k,i,j 1 k 2 θ min,,off,j θ max,off,j i off,h,j s ( θ bk ) p n=j ( θ in )sin θ in d θ in θ min,,off,j θ max,off,j p n=j ( θ in )sin θ in d θ in exp( 2 q=k i σ ext,off ( z q , z k , z i , t j )Δ z q ) ] .
θ bk =π( θ in ( t j ) θ sat ),
θ sat = tan 1 [ ( z i z i1 )tan( θ in ( t j ))/ ( z sat + z i ) ].
i w,h,j s ( θ )= | E w,h,j s | 2 ϕ / ( | E w, x, j in | 2 + | E w, y, j in | 2 ) .
E w,j s = L ϕ D s n=j L ϕ E w,j in .
S n =[ S 11,n ( θ ) S 12,n ( θ ) 0 0 S 12,n ( θ ) S 22,n ( θ ) 0 0 0 0 S 33,n ( θ ) S 34,n ( θ ) 0 0 S 34,n ( θ ) S 44,n ( θ ) ]
i w, x,j s ( θ )= ( m 11,n + m 22,n ) | E w, x,j in | 2 +( m 11,n m 22,n ) | E w, y,j in | 2 | E w, j in | 2 = ( m 11,n + m 22,n )+( m 11,n m 22,n ) Q w, j in 1+ Q w, j in (n=j),
i w, y,j s ( θ )= ( m 11,n m 22,n ) | E w, x,j in | 2 +( m 11,n + m 22,n ) | E w, y,j in | 2 | E w, j in | 2 = ( m 11,n m 22,n )+( m 11,n + m 22,n ) Q w, j in 1+ Q w, j in (n=j),
Q w, j in = | E w, y,j in | 2 / | E w, x, j in | 2 , ( w=ss,on,off )
m 11,n + m 22,n =A( θ )[ 2 S 11,n ( θ )+ S 22,n ( θ ) ]+3B( θ ) S 12,n ( θ )2C( θ ) S 33,n ( θ ),
m 11,n m 22,n =A( θ )[ 2 S 11,n ( θ ) S 22,n ( θ ) ]+B( θ ) S 12,n ( θ )+2C( θ ) S 33,n ( θ ).
A( θ )= ( 1+ cos 2 θ )/8 ,
B( θ )= sin 2 θ/8 ,
C( θ )= cosθ/8 .
Δ j+1, j = [ m 11,n ( θ ) m 22,n ( θ ) ]+[ m 11,n ( θ )+ m 22,n ( θ ) ] Δ j, j [ m 11,n ( θ )+ m 22,n ( θ ) ]+[ m 11,n ( θ ) m 22,n ( θ ) ] Δ j, j ( j 1j1,n=j ),
θ ref = tan 1 [ r ref,i / l i ](i= j ),
r ref,i =min[ 0.5 r fov ( z i ), k=0 i1 ( l k+1 1 g n=k 2 ) ](i= j ).
Q on, j in = Δ j, j δ j j ,
θ max,on,j = tan 1 [ r fov ( z i )/ l i ](i=j),
θ min,on,j ={ θ df ( θ max,on,j > θ df ) 0( θ max,on,j θ df ) .
Q off,j+1 in = 0 π { ( m 11,n m 22,n )+( m 11,n + m 22,n ) Q off,j in } p n ( θ )sinθdθ 0 π { ( m 11,n + m 22,n )+( m 11,n m 22,n ) Q off,j in } p n ( θ )sinθdθ ( j 1j1,n=j ),
θ off,ki ( t j )= cos 1 [ ( z i z k )/ { ( t j+1/2 c z k ) } ].
θ min,off,j = θ off,ki ( t j1 ),
θ max,off,j = θ off,ki ( t j ).
δ( t j )= I tot,y ( t j ) I tot,x ( t j ) = I ss,y ( t j )+ I ms,tot,on,y ( t j )+ i=1 j I ms,off,y ( z i , t j ) I ss,x ( t j )+ I ms,tot,on,x ( t j )+ i=1 j I ms,off,x ( z i , t j ) .
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.