Skip to main content

The LIDORT and VLIDORT Linearized Scalar and Vector Discrete Ordinate Radiative Transfer Models: Updates in the Last 10 Years

  • Chapter
  • First Online:
Springer Series in Light Scattering

Part of the book series: Springer Series in Light Scattering ((SSLS))

Abstract

It has been 10 years since the last major review paper on the LIDORT and VLIDORT radiative transfer models; this paper appeared in Light Scattering Reviews, Volume 3 (Spurr in Light scattering reviews. Springer, Berlin, 2008), hereinafter referenced as [R1]).

This is a preview of subscription content, log in via an institution to check access.

Access this chapter

Chapter
USD 29.95
Price excludes VAT (USA)
  • Available as PDF
  • Read on any device
  • Instant download
  • Own it forever
eBook
USD 109.00
Price excludes VAT (USA)
  • Available as EPUB and PDF
  • Read on any device
  • Instant download
  • Own it forever
Hardcover Book
USD 139.99
Price excludes VAT (USA)
  • Durable hardcover edition
  • Dispatched in 3 to 5 business days
  • Free shipping worldwide - see info

Tax calculation will be finalised at checkout

Purchases are for personal use only

Institutional subscriptions

References

  • Anderson E, Bai Z, Bischof C, Demmel J, Dongarra J, Du Croz J, Greenbaum A, Hammarling S, McKenney A, Ostrouchov A, Sorensen D (1995) LAPACK User’s Guide, 2nd edn. Philadephia, Society for Industrial and Applied Mathematics

    Google Scholar 

  • Chandrasekhar S (1960) Radiative transfer. Dover Publications Inc., New York

    Google Scholar 

  • Chami M, Santer R, Dilligeard E (2001) Radiative transfer model for the computation of radiance and polarization in an ocean-atmosphere system: polarization properties of suspended matter for remote sensing. Appl Opt 40:2398–2416

    Article  ADS  Google Scholar 

  • Coulson K, Dave J, Sekera D (1960) Tables related to radiation emerging from planetary atmosphere with Rayleigh scattering. University of California Press, Berkeley

    Google Scholar 

  • Cox C, Munk W (1954a) Statistics of the sea surface derived from sun glitter. J Mar Res 13:198–227

    Google Scholar 

  • Cox C, Munk W (1954b) Measurement of the roughness of the sea surface from photographs of the sun’s glitter. J Opt Soc Am 44:838–850

    Article  ADS  Google Scholar 

  • de Rooij WA, Van der Stap CCAH (1984) Expansion of Mie scattering matrices in generalized spherical functions, Astron Astrophys 131:237–248

    Google Scholar 

  • Doicu A, Trautmann T (2009) Two linearization methods for atmospheric remote sensing. J Quant Spectrosc Radiat Transf 110:477–490

    Article  ADS  Google Scholar 

  • Efremenko D, Doicu A, Loyola D, Trautmann T (2013) Acceleration techniques for the discrete ordinate method. J Quant Spectrosc Radiat Transf 114:73–81

    Article  ADS  Google Scholar 

  • Frankenberg C, O’Dell C, Guanter L, McDuffie J (2012) Remote sensing of near-infrared chlorophyll fluorescence from space in scattering atmospheres: implications for its retrieval and interferences with atmospheric CO2 retrievals. Atmos Meas Tech 5:2081–2094

    Article  Google Scholar 

  • Garcia RDM, Siewert CE (1989) The FN method for radiative transfer models that include polarization. J Quant Spectrosc Radiat Transf 41:117–145

    Article  ADS  Google Scholar 

  • Hapke B (1993) Theory of reflectance and emittance spectroscopy. Cambridge University Press, Cambridge, UK

    Book  Google Scholar 

  • Hasekamp OP, Landgraf J (2002) A linearized vector radiative transfer model for atmospheric trace gas retrieval. J Quant Spectrosc Radiat Transf 75:221–238

    Article  ADS  Google Scholar 

  • Hovenier JW, Van der Mee C, Domke H (2004) Transfer of polarized light in planetary atmospheres: Basic concepts and practical methods, Kluwer, Dordrecht

    Google Scholar 

  • Huang H, Qin W, Spurr RJ, Liu Q (2017) Evaluation of atmospheric effects on land surface directional reflectance with the coupled RAPID and VLIDORT models. Geosci Remote Sens Lett 14:916–920

    Article  ADS  Google Scholar 

  • Jin Z, Charlock T, Rutledge K, Stamnes K, Wang Y (2006) Analytic solution of radiative transfer in the coupled atmosphere-ocean system with a rough surface. Appl Opt 45:7433–7455

    Article  ADS  Google Scholar 

  • Kotchenova SY, Vermote EF, Matarrese R, Klemm FJ (2006) Validation of a vector version of the 6S radiative transfer code for atmospheric correction of satellite data. Part I: Path Radiance Appl Opt 45:6762–6774

    ADS  Google Scholar 

  • Lucht W, Schaaf C, Strahler A (2000) An algorithm for the retrieval of Albedo from space using semi-empirical BRDF models. IEEE Trans Geosci Remote Sens 38:977

    Article  ADS  Google Scholar 

  • Lucht W, Roujean J-L (2000) Considerations in the Parametric modeling of BRDF and Albedo from multiangular satellite sensor observations. Remote Sens Rev 18:343–379

    Article  Google Scholar 

  • Lyapustin A, Gatebe CK, Kahn R, Brandt R, Redemann J, Russell P, King MD, Pedersen CA, Gerland S, Poudyal R, Marshak A (2010) Analysis of snow bidirectional reflectance from ARCTAS Spring-2008 Campaign. Atmos Chem Phys 10:4359–4375

    Article  ADS  Google Scholar 

  • Mackowski DW, Mishchenko MI (1996) Calculation of the T matrix and the scattering matrix for ensembles of spheres, J Opt Soc Am A 13:2266–2278

    Article  ADS  Google Scholar 

  • Maignan F, Bréon F-M, Fédèle E, Bouvier M (2009) Polarized reflectance of natural surfaces: Spaceborne measurements and analytical modeling. Rem Sens Environ 113:2642–2650

    Article  ADS  Google Scholar 

  • Meng Z, Yang P, Kattawar GW, Bi L, Liou K-N, Lazslo I (2010) Single-scattering properties of tri-axial ellipsoidal mineral dust aerosols: A database for application to radiative transfer calculations. J Aer Sci 41:501–512

    Article  ADS  Google Scholar 

  • Mishchenko MI, Travis LD (1998) Capabilities and limitations of a current FORTRAN implementation of the T-matrix method for randomly oriented, rotationally symmetric scatterers, JQSRT 60:309–324

    Article  ADS  Google Scholar 

  • Mishchenko MI, Travis LD (1998) Satellite retrieval of aerosol properties over the ocean using polarization as well as intensity of reflected sunlight. J Geophys Res 102:16989

    Article  Google Scholar 

  • Mishchenko MI, Travis LD, Lacis LL (2006) Scattering, absorption and emission of light by small particles. Cambridge University Press, Cambridge, U.K.

    Google Scholar 

  • Mishchenko MI (2014) Electromagnetic scattering by particles and particle groups: an introduction. Cambridge University Press, ISBN 9780521519922

    Google Scholar 

  • Morel A, Gentili B (2009) A simple band ratio technique to quantify the colored dissolved and detrital organic material from ocean color remotely sensed data. Remote Sens Environ 113

    Article  ADS  Google Scholar 

  • Nakajima T, Tanaka M (1988) Algorithms for radiative intensity calculations in moderately thick atmospheres using a truncation approximation. J Quant Spectrosc Radiat Transf 40:51–69

    Article  ADS  Google Scholar 

  • Natraj V, Shia RL, Yung YL (2010) On the use of principal component analysis to speed up radiative transfer calculations. J Quant Spectrosc Radiat Transfer 111:810–816, https://doi.org/10.1016/j.jqsrt.2009.11.004

    Article  ADS  Google Scholar 

  • Natraj V, Hovenier JW (2012) Polarized light reflected and transmitted by thick rayleigh scattering atmospheres. Astrophys J 748:28

    Article  ADS  Google Scholar 

  • Rahman H, Pinty B, Verstrate M (1993) Coupled surface-atmospheric reflectance (CSAR) model. 2. Semi-empirical surface model usable with NOAA advanced very high resolution radiometer data. J Geophys Res 98:20791

    Article  ADS  Google Scholar 

  • Rodgers CD (2000) Inverse Methods for Atmospheric Sounding: Theory and Practice, World Scientific Publishing Co. Pte. Ltd., Singapore

    Book  Google Scholar 

  • Rozanov V, Rozanov A (2007) Relationship between different approaches to derive weighting functions related to atmospheric remote sensing problems. J Quant Spectrosc Radiat Transf 105(2):217–242

    Article  ADS  Google Scholar 

  • Sancer M (1969) Shadow-corrected electromagnetic scattering from a randomly-rough ocean surface. IEEE Trans Antennas Propag AP-17:557–585

    Google Scholar 

  • Sayer A, Thomas G, Grainger R (2010) A sea-surface reflectance model for (A)ATSR, and application to aerosol retrievals. Atmos Meas Tech 3:813–838

    Article  ADS  Google Scholar 

  • Siewert CE (1982) On the phase matrix basic to the scattering of polarized light. Astron Astrophys 109:195–200

    Google Scholar 

  • Siewert CE (2000a) A concise and accurate solution to Chandrasekhar’s basic problem in radiative transfer. J Quant Spectrosc Radiat Transf 64:109–130

    Article  ADS  Google Scholar 

  • Siewert CE (2000b) A discrete-ordinates solution for radiative transfer models that include polarization effects. J Quant Spectrosc Radiat Transf 64:227–254

    Article  ADS  Google Scholar 

  • Spurr R (2002) Simultaneous derivation of intensities and weighting functions in a general pseudo-spherical discrete ordinate radiative transfer treatment. J Quant Spectrosc Radiat Transf 75:129–175

    Article  ADS  Google Scholar 

  • Spurr RJD (2004) A New approach to the retrieval of surface properties from earthshine measurements. J Quant Spectrosc Radiat Transf 83:15–46

    Article  ADS  Google Scholar 

  • Spurr RJD (2006) VLIDORT: a linearized pseudo-spherical vector discrete ordinate radiative transfer code for forward model and retrieval studies in multilayer multiple scattering media. J Quant Spectrosc Radiat Transf 102(2):316–342. https://doi.org/10.1016/j/jqsrt.2006.05.005

    Article  ADS  Google Scholar 

  • Spurr R (2008) LIDORT and VLIDORT: Linearized pseudo-spherical scalar and vector discrete ordinate radiative transfer models for use in remote sensing retrieval problems. In: Kokhanovsky A (ed) Light scatter reviews, vol 3. Springer, Berlin

    Google Scholar 

  • Spurr R, Natraj V (2011) A linearized two-stream radiative transfer code for fast approximation of multiple-scatter fields. J Quant Spectrosc Radiat Transfer 112:2630–2637

    Article  ADS  Google Scholar 

  • Spurr R, Christi M (2014) On the generation of atmospheric property Jacobians from the (V)LIDORT linearized radiative transfer models. J Quant Spectrosc Radiat Transf. https://doi.org/10.1016/j.jqsrt.2014.03.011

    Article  ADS  Google Scholar 

  • Spurr R, Christi M (2018) LIDORT-RRS: a linearized discrete ordinate radiative transfer model with first order rotational raman scattering, paper in preparation

    Google Scholar 

  • Spurr R, Kurosu T, Chance K (2001) A linearized discrete ordinate radiative transfer model for atmospheric remote sensing retrieval. J Quant Spectrosc Radiat Transf 68:689–735

    Article  ADS  Google Scholar 

  • Spurr RJD, de Haan J, van Oss R, Vasilkov A (2008) Discrete ordinate theory in a stratified medium with first order rotational raman scattering; a general quasi-analytic solution. J Quant Spectrosc Radiat Transfer 109:404–425. https://doi.org/10.1016/j.jqsrt.2007.08.011

    Article  ADS  Google Scholar 

  • Spurr R, Wang J, Zeng J, Mishchenko M (2012) Linearized T-Matrix and Mie scattering computations. J Quant Spectrosc Radiat Transfer 113:425–439

    Article  ADS  Google Scholar 

  • Spurr R, Natraj V, Kopparla P, Christi M (2016) Application of principal component analysis (PCA) to performance enhancement of hyperspectral radiative transfer computations, in “Principal Component Analysis: Methods, Applications and Technology”, NOVA publishers

    Google Scholar 

  • Stamnes K, Conklin P (1984) A new multi-layer discrete ordinate approach to radiative transfer in vertically inhomogeneous atmospheres. J Quant Spectrosc Radiat Transfer 31:273

    Article  ADS  Google Scholar 

  • Stamnes K, Tsay S-C, Nakajima T (1988) Computation of eigenvalues and eigenvectors for discrete ordinate and matrix operator method radiative transfer. J Quant Spectrosc Radiat Transf 39:415–419

    Google Scholar 

  • Ustinov EA (2005) Atmospheric weighting functions and surface partial derivatives for remote sensing of scattering planetary atmospheres in thermal spectral region: General adjoint approach. J Quant Spectrosc Radiat Transf 92:351–371

    Article  ADS  Google Scholar 

  • Van Oss RF, Spurr RJD (2002) Fast and accurate 4 and 6 stream linearized discrete ordinate radiative transfer models for ozone profile retrieval. J Quant Spectrosc Radiat Transf 75:177–220

    Article  ADS  Google Scholar 

  • Vermote EF, Tanré D, Deuzé JL, Herman M, Morcrette JJ (1997) Second simulation of the satellite signal in the solar spectrum, 6S: an overview. IEEE Trans Geosci Remote Sens 35:675–686

    Article  ADS  Google Scholar 

  • Vestrucci M, Siewert CE (1984) A numerical evaluation of an analytical representation of the components in a Fourier decomposition of the phase matrix for the scattering of polarized light. JQSRT 31:177–183

    Article  ADS  Google Scholar 

  • Wanner W, Li X, Strahler A (1995) On the derivation of kernels for kernel-driven models of bidirectional reflectance. J Geophys Res 100:21077

    Article  ADS  Google Scholar 

  • Wiscombe W (1977) The delta-M method: rapid yet accurate radiative flux calculations for strongly asymmetric phase function

    Google Scholar 

Download references

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Robert Spurr .

Editor information

Editors and Affiliations

Appendices

Appendix A. RTE Solutions for VLIDORT

1.1 A.1 Azimuthal Separation

In applications involving randomly oriented particles with one plane of symmetry, the scattering matrix \( {\mathbf{F}}\left(\Theta \right) \) in Eq. (2.5) has six independent entries:

$$ {\mathbf{F}}\left(\Theta \right) = \left( {\begin{array}{*{20}c} {a_{1} \left(\Theta \right)} & {b_{1} \left(\Theta \right)} & 0 & 0 \\ {b_{1} \left(\Theta \right)} & {a_{2} \left(\Theta \right)} & 0 & 0 \\ 0 & 0 & {a_{3} \left(\Theta \right)} & {b_{2} \left(\Theta \right)} \\ 0 & 0 & { - b_{2} \left(\Theta \right)} & {a_{4} \left(\Theta \right)} \\ \end{array} } \right). $$
(A.1)

For this form of the scattering matrix, one can develop expansions of these functions in terms of a set of generalized spherical functions \( P_{mn}^{l} \left( {\cos\Theta } \right) \) (Mishchenko et al. 2006):

$$ \begin{aligned} a_{1} \left(\Theta \right) & = \sum\limits_{l = 0}^{M} {\beta_{l} P_{00}^{l} \left( {\cos\Theta } \right)} ; \\ a_{2} \left(\Theta \right) + a_{3} \left(\Theta \right) & = \sum\limits_{l = 0}^{M} {\left( {\alpha_{l} + \zeta_{l} } \right)P_{2,2}^{l} \left( {\cos\Theta } \right)} ; \\ a_{2} \left(\Theta \right) - a_{3} \left(\Theta \right) & = \sum\limits_{l = 0}^{M} {\left( {\alpha_{l} - \zeta_{l} } \right)P_{2, - 2}^{l} \left( {\cos\Theta } \right)} ; \\ a_{4} \left(\Theta \right) & = \sum\limits_{l = 0}^{M} {\delta_{l} P_{00}^{l} \left( {\cos\Theta } \right)} \\ \end{aligned} $$
(A.2)
$$ b_{1} \left(\Theta \right) = \sum\limits_{l = 0}^{M} {\gamma_{l} P_{02}^{l} \left( {\cos\Theta } \right)} ; \quad b_{2} \left(\Theta \right) = - \sum\limits_{l = 0}^{M} {\epsilon_{l} P_{02}^{l} \left( {\cos\Theta } \right)} $$
(A.3)

The (1,1) entry in \( {\mathbf{F}}\left(\Theta \right) \) is just the phase function and satisfies the normalization condition:

$$ \frac{1}{2}\int\limits_{0}^{\pi } {} a_{1} \left(\Theta \right){ \sin }\Theta d\Theta = 1. $$
(A.4)

The set of six expansion coefficients (“Greek constants”) \( \left\{ {\alpha_{l} ,\beta_{l} ,\gamma_{l} ,\delta_{l} ,\varepsilon_{l} ,\zeta_{l} } \right\} \) are key inputs to VLIDORT, and the number of terms \( M \) in these expansions depends on the desired level of numerical accuracy. Here, \( \left\{ {\beta_{l} } \right\} \) are the phase function expansion coefficients as used in the scalar RTE. These “Greek constants” specify the polarized-light single-scattering law, and there are a number of efficient analytical techniques for their computation, not only for spherical particles (see for example de Rooij and van der Stap 1984) but also for randomly-oriented homogeneous and inhomogeneous non-spherical particles and aggregated scatterers (Hovenier et al. 2004; Mackowski and Mishchenko 1996; Mishchenko and Travis 1998).

To proceed with the RTE solution, it is necessary to make Fourier decompositions (in terms of the cosine and sine of the relative azimuth angle between incident and scattered light directions) of the phase matrix and the Stokes vector in order to separate the azimuthal dependence.

$$ {\mathbf{I}}\left( {x,\mu ,\phi - \phi^{{\prime }} } \right) = \frac{1}{2}\sum\limits_{m = 0}^{M} {\left( {2 - \delta_{m,0} } \right){\varvec{\Phi}}_{m} \left( {\phi - \phi^{{\prime }} } \right){\mathbf{I}}_{m} \left( {x,\mu } \right)} ; $$
(A.5)
$$ {\varvec{\Pi}}\left( {\mu ,\phi ,\mu^{{\prime }} ,\phi^{{\prime }} } \right) = \frac{1}{2}\sum\limits_{m = 0}^{M} {\left( {2 - \delta_{m,0} } \right)\left[ {{\mathbf{C}}_{m} \left( {\mu ,\mu^{{\prime }} } \right){ \cos } m\left( {\phi - \phi^{{\prime }} } \right) + {\mathbf{S}}_{m} \left( {\mu ,\mu^{{\prime }} } \right){ \sin } m\left( {\phi - \phi^{{\prime }} } \right)} \right]} ; $$
(A.6)

Here, \( {\varvec{\Phi}}_{m} \left( \phi \right) \equiv Diag\left[ {\cos m\phi ,\cos m\phi ,\sin m\phi ,\sin m\phi } \right] \). We follow the formulation for azimuthal separation of the scattering matrix developed by Siewert (1982), Vestrucci and Siewert (1984). Most vector radiative transfer models now follow this work. Accordingly, we write:

$$ {\mathbf{C}}_{m} \left( {\mu ,\mu^{{\prime }} } \right) = {\mathbf{A}}_{m} \left( {\mu ,\mu^{{\prime }} } \right) + {\mathbf{DA}}_{m} \left( {\mu ,\mu^{{\prime }} } \right){\mathbf{D}};\, {\mathbf{S}}_{m} \left( {\mu ,\mu^{{\prime }} } \right) = {\mathbf{A}}_{m} \left( {\mu ,\mu^{{\prime }} } \right){\mathbf{D}} - {\mathbf{DA}}_{m} \left( {\mu ,\mu^{{\prime }} } \right); $$
(A.7)
$$ {\mathbf{A}}_{m} \left( {\mu ,\mu^{{\prime }} } \right) = \sum\limits_{l = m}^{M} {{\mathbf{P}}_{l}^{m} \left( \mu \right){\mathbf{B}}_{l} {\mathbf{P}}_{l}^{m} \left( {\mu^{{\prime }} } \right)} ;\,{\mathbf{D}} = Diag\left[ {1,1, - 1, - 1} \right];\varvec{ } $$
(A.8)
$$ {\mathbf{B}}_{l} = \left( {\begin{array}{*{20}c} {\beta_{l} } & {\gamma_{l} } & 0 & 0 \\ {\gamma_{l} } & {\alpha_{l} } & 0 & 0 \\ 0 & 0 & {\delta_{l} } & { - \varepsilon_{l} } \\ 0 & 0 & {\varepsilon_{l} } & {\zeta_{l} } \\ \end{array} } \right); {\mathbf{P}}_{l}^{m} \left( \mu \right) = \left( {\begin{array}{*{20}c} {P_{l}^{m} \left( \mu \right)} & {\gamma_{l} } & 0 & 0 \\ {\gamma_{l} } & {R_{l}^{m} \left( \mu \right)} & { - T_{l}^{m} \left( \mu \right)} & 0 \\ 0 & { - T_{l}^{m} \left( \mu \right)} & {R_{l}^{m} \left( \mu \right)} & 0 \\ 0 & 0 & 0 & {P_{l}^{m} \left( \mu \right)} \\ \end{array} } \right). $$
(A.9)

The “Greek matrices” contain the spherical-function expansion coefficients, while the matrices \( {\mathbf{P}}_{l}^{m} \left( \mu \right) \) contain the associated Legendre functions \( P_{l}^{m} \left( \mu \right) \) and the functions \( R_{l}^{m} \left( \mu \right) \) and \( T_{l}^{m} \left( \mu \right) \) which are closely related to the generalized spherical functions \( P_{mn}^{l} \left( \mu \right) \) (for details, see for example Siewert 2000b).

This azimuth separation process yields the following RTE for the Fourier component \( {\mathbf{I}}_{m} \left( {x,\mu } \right) \):

$$ \mu \frac{{d{\mathbf{I}}_{m} \left( {x,\mu } \right)}}{dx} = - {\mathbf{I}}_{m} \left( {x,\mu } \right) + \frac{\omega }{2}\sum\limits_{l = m}^{M} {{\mathbf{P}}_{l}^{m} \left( \mu \right){\mathbf{B}}_{l} } \int\limits_{ - 1}^{1} {{\mathbf{P}}_{l}^{m} \left( {\mu^{{\prime }} } \right){\mathbf{I}}_{m} \left( {x,\mu^{{\prime }} } \right)d\mu^{{\prime }} + {\mathbf{Q}}_{m} \left( {x,\mu } \right)} . $$
(A.10)

For the solar source term, with solar direction \( \left\{ { - \mu_{0} ,\phi_{0} } \right\} \), we have.

$$ {\mathbf{Q}}_{m}^{ \odot } \left( {x,\mu } \right) = \frac{\omega }{2}\sum\limits_{l = m}^{M} {{\mathbf{P}}_{l}^{m} \left( \mu \right){\mathbf{B}}_{l} {\mathbf{P}}_{l}^{m} \left( { - \mu_{0} } \right){\mathbf{F}}_{ \odot } { \exp }\left[ { - \tau_{ \odot } \left( {x,\mu } \right)} \right],} $$
(A.11)

in terms of the TOA solar flux \( {\mathbf{F}}_{ \odot } = \left[ {F_{ \odot } ,0,0,0} \right]^{\text{T}} \) and solar beam attenuation \( { \exp }\left[ { - \tau_{ \odot } \left( {x,\mu_{0} } \right)} \right] \), where \( \tau_{ \odot } \left( {x,\mu_{0} } \right) \) is the beam optical depth in a spherical-shell atmosphere.

1.2 A.2 Homogeneous Solutions in VLIDORT

1.2.1 A.2.1 Eigenvalue Solutions

In the discrete-ordinate solution method, we solve for each Fourier component in Eq. (A.10) by first finding the homogeneous solutions (without the solar or thermal source terms). With the familiar discrete ordinate quadrature with stream directions and weights \( \left\{ { \pm \mu_{i} ,c_{i} } \right\} \) for \( i = 1, \ldots N_{d} \), where \( N_{d} \) is the number of discrete ordinates in the polar half space. The resulting vector RTE for these streams is then

$$ \begin{aligned} \pm \mu_{i} \frac{{d{\mathbf{I}}_{m} \left( {x, \pm \mu_{i} } \right)}}{dx} \pm {\mathbf{I}}_{m} \left( {x, \pm \mu_{i} } \right) & = \frac{\omega }{2}\sum\limits_{l = m}^{M} {{\mathbf{P}}_{l}^{m} \left( { \pm \mu_{i} } \right){\mathbf{B}}_{l} } \sum\limits_{j = 1}^{{N_{d} }} {\left[ {{\mathbf{P}}_{l}^{m} \left( {\mu_{j} } \right){\mathbf{I}}_{m} \left( {x,\mu_{j} } \right)} \right.} \\ & \quad + \left. {{\mathbf{P}}_{l}^{m} \left( { - \mu_{j} } \right){\mathbf{I}}_{m} \left( {x, - \mu_{j} } \right)} \right] \\ \end{aligned} $$
(A.12)

There are \( 8N_{d} \) coupled first-order linear differential equations for \( {\mathbf{I}}_{m} \left( {x, \pm \mu_{i} } \right) \) in this system, which is solved by eigenvalue methods, using the ansatz:

$$ {\mathbf{I}}_{\alpha } \left( {x, \pm \mu_{i} } \right) = {\mathbf{W}}_{\alpha } \left( { \pm \mu_{i} } \right)\exp \left[ { - k_{\alpha } x} \right]. $$
(A.13)

We then define the following two vectors of rank \( 4N_{d} \):

$$ {\mathbb{W}}_{\alpha }^{ \pm } = \left[ {{\mathbf{W}}_{\alpha }^{\text{T}} \left( { \pm \mu_{1} } \right),{\mathbf{W}}_{\alpha }^{\text{T}} \left( { \pm \mu_{2} } \right), \ldots ,{\mathbf{W}}_{\alpha }^{\text{T}} \left( { \pm \mu_{{N_{d} }} } \right)} \right]^{\text{T}} . $$
(A.14)

The system is decoupled using sum and difference vectors \( {\mathbb{X}}_{\alpha } = {\mathbb{W}}_{\alpha }^{ + } + {\mathbb{W}}_{\alpha }^{ - } \) and \( {\mathbb{Y}}_{\alpha } = {\mathbb{W}}_{\alpha }^{ + } - {\mathbb{W}}_{\alpha }^{ - } \), and the order is then be reduced from \( 8N_{d} \) to \( 4N_{d} \). The result is an eigenproblem for the collection of separation constants and associated solution vectors \( \left\{ {k_{\alpha } ,{\mathbb{X}}_{\alpha } } \right\} \), where index \( \alpha = 1, \ldots ,4N_{d} \) labels the eigensolutions. The eigenmatrix for this system is constructed from the optical property inputs \( \left\{ {\omega ,{\mathbf{B}}_{l} } \right\} \) and combination products of matrices \( {\mathbf{P}}_{l}^{m} \left( { \pm \mu_{i} } \right) \). The eigenproblem is Siewert (2000b):

$$ {\mathbb{X}}_{\alpha }^{*} {\mathbb{G}} = k_{\alpha }^{2} {\mathbb{X}}_{\alpha }^{*} ; {{\mathbb{G}}{\mathbb{X}}}_{\alpha } = k_{\alpha }^{2} {\mathbb{X}}_{\alpha } ; $$
(A.15a)
$$ {\mathbb{G}} = {\mathbb{F}}^{ + } {\mathbb{F}}^{ - } ; {\mathbb{F}}^{ \pm } = \left[ {{\mathbb{E}} - \frac{\omega }{2}\sum\limits_{l = m}^{M} {{\mathbb{P}}\left( {l,m} \right){\mathbf{B}}_{l} {\mathbf{A}}^{ \pm } {\mathbb{P}}^{\text{T}} \left( {l,m} \right){\mathbb{C}}} } \right]{\mathbb{M}}^{ - 1} ; $$
(A.15b)
$$ {\mathbb{P}}\left( {l,m} \right) = Diag\left[ {{\mathbf{P}}_{l}^{m} \left( {\mu_{1} } \right),{\mathbf{P}}_{l}^{m} \left( {\mu_{2} } \right), \ldots ,{\mathbf{P}}_{l}^{m} \left( {\mu_{{N_{d} }} } \right)} \right]^{\text{T}} ; $$
(A.15c)
$$ {\mathbb{M}} = Diag\left[ {\mu_{1} {\mathbf{E}},\mu_{2} {\mathbf{E}}, \ldots ,\mu_{{N_{d} }} {\mathbf{E}}} \right];\varvec{ }{\mathbb{C}}\varvec{ } = Diag\left[ {c_{1} {\mathbf{E}},c_{2} {\mathbf{E}}, \ldots ,c_{{N_{d} }} {\mathbf{E}}} \right]. $$
(A.15d)

Here, \( \varvec{ }{\mathbf{A}}^{ \pm } = {\mathbf{E}} \pm \left( { - 1} \right)^{l - m} {\mathbf{D}};{\mathbf{E}} = Diag\left[ {1,1,1,1} \right];\varvec{ }{\mathbf{D}} = Diag\left[ {1,1, - 1, - 1} \right] \), \( {\mathbb{E}} \) is the \( 4N_{d} \times 4N_{d} \) identity matrix; \( {\mathbb{X}}_{\alpha }^{*} \) and \( {\mathbb{X}}_{\alpha } \) are the left and right eigenvectors respectively, with \( {\mathbb{X}}_{\alpha }^{*} \) the conjugate transpose of \( {\mathbb{X}}_{\alpha } \). The link between \( {\mathbb{X}}_{\alpha } \) and solution vectors \( {\mathbb{W}}_{\alpha }^{ \pm } \) comes through the auxiliary equations:

$$ {\mathbb{W}}_{\alpha }^{ \pm } = \frac{1}{2}{\mathbb{M}}^{ - 1} \left[ {{\mathbb{E}} \pm \frac{1}{{k_{\alpha } }}{\mathbb{F}}^{ + } } \right]{\mathbb{X}}_{\alpha } . $$
(A.16)

Eigenvalues occur in pairs \( \left\{ { \pm k_{\alpha } } \right\} \). Left and right eigenvectors share the same spectrum of eigenvalues. As noted by Siewert (2000b), both complex- and real-variable eigensolutions may be present in the full Stokes 4-vector case (rank \( 4N_{d} \)). Eigensolutions may be determined numerically with the complex-variable eigensolver DGEEV from the LAPACK suite (Anderson et al. 1995). DGEEV generates both right and left eigenvectors, which have unit moduli. In the scalar case (no polarization, solutions only for the I-component of the Stokes vector), the eigensystem has rank \( N_{d} \), the eigenmatrix is symmetric and all eigensolutions are real-valued. In this case, the eigensolver module ASYMTX (Stamnes et al. 1988) may be used. ASYMTX is an adaptation of the LAPACK routine for real-valued problems; it delivers only the right eigenvectors. For vector solutions neglecting circular polarization (Stokes I, Q and U only), complex eigensolutions are absent, and one may then use the faster ASYMTX routine.

Returning to the full Stokes 4-vector case, the complete homogeneous solution in one layer is then:

$$ {\mathbb{I}}^{ + } \left( x \right) = {\mathbb{D}}^{ + } \sum\limits_{\alpha = 1}^{{4N_{d} }} {\left\{ {{\mathfrak{L}}_{\alpha } {\mathbb{W}}_{\alpha }^{ + } \exp \left[ { - k_{\alpha } x} \right] + {\mathfrak{M}}_{\alpha } {\mathbb{W}}_{\alpha }^{ - } \exp \left[ { - k_{\alpha } \left( {\Delta - x} \right)} \right]} \right\}} ; $$
(A.17a)
$$ {\mathbb{I}}^{ - } \left( x \right) = {\mathbb{D}}^{ - } \sum\limits_{\alpha = 1}^{{4N_{d} }} {\left\{ {{\mathfrak{L}}_{\alpha } {\mathbb{W}}_{\alpha }^{ - } \exp \left[ { - k_{\alpha } x} \right] + {\mathfrak{M}}_{\alpha } {\mathbb{W}}_{\alpha }^{ + } \exp \left[ { - k_{\alpha } \left( {{\Delta } - x} \right)} \right]} \right\}} ; $$
(A.17b)

Here, \( {\mathbb{D}}^{ - } = Diag\left[ {{\mathbf{D}},{\mathbf{D}}, \ldots {\mathbf{D}}} \right] \), and \( {\mathbb{D}}^{ + } = {\mathbb{E}} \); these matrices arise from application of symmetry relations (Siewert 2000b). The use of optical thickness \( \Delta - x \) in the second exponential ensures that solutions remain bounded (Stamnes and Conklin 1984). The quantities \( \left\{ {{\mathfrak{L}}_{\alpha } ,{\mathfrak{M}}_{\alpha } } \right\} \) are the constants of integration; determined by application of the boundary conditions and solution of the resulting boundary-value problem.

In the Stokes 4-vector case, some contributions to \( {\mathbb{I}}^{ \pm } \left( x \right) \) will be complex, some real. It is understood that we compute the real parts of any complex variable expressions. Thus for example if \( \left\{ {k_{\alpha } ,{\mathbb{W}}_{\alpha }^{ - } } \right\} \) is a complex eigensolution with associated (complex) integration constant \( {\mathfrak{L}}_{\alpha } \), the real part of the solution will be:

$$ {\text{Re}}\left[ {{\mathfrak{L}}_{\alpha } {\mathbb{W}}_{\alpha }^{ - } e^{{ - k_{\alpha } x}} } \right] = {\text{Re}}\left[ {{\mathfrak{L}}_{\alpha } } \right]{\text{Re}}\left[ {{\mathbb{W}}_{\alpha }^{ - } e^{{ - k_{\alpha } x}} } \right] - {\text{Im}}\left[ {{\mathfrak{L}}_{\alpha } } \right]{\text{Im}}\left[ {{\mathbb{W}}_{\alpha }^{ - } e^{{ - k_{\alpha } x}} } \right]. $$
(A.18)

From a bookkeeping standpoint, one must keep count of the number of real and complex solutions, and treat them separately in the numerical implementation. For clarity of exposition, we have not made an explicit separation of complex variables, and it will be clear from the context whether real or complex variables are under consideration.

1.2.2 A.2.2 Linearization of the Eigenvalue Solutions

For a single layer, we require derivatives of the eigensolution \( \left\{ {k_{\alpha } ,{\mathbb{W}}_{\alpha }^{ \pm } } \right\} \) with respect to some atmospheric variable \( \xi \) in that layer. The starting point for the differentiation is the set of linearized optical properties \( {\mathcal{V}} \equiv \xi \partial \Delta /\partial \xi ;{ \mathcal{U}} \equiv \xi \partial\upomega/\partial \xi ; {\boldsymbol{{\mathcal{Z}}}}_{l} \equiv \xi \partial {\mathbf{B}}_{l} /\partial \xi \), that is, normalized partial derivatives of the set of IOPs \( \left\{ {\Delta ,\upomega, {\mathbf{B}}_{l} } \right\} \). The eigensolution depends only on the quantities \( \left\{ {\upomega, {\mathbf{B}}_{l} } \right\} \). Applying the linearization operator \( {\mathcal{L}} \equiv \xi \partial /\partial \xi \) to the eigenmatrix, we find

$$ {\mathcal{L}}\left( {\mathbb{G}} \right) = { \mathcal{L}}\left( {{\mathbb{F}}^{ + } } \right){\mathbb{F}}^{ - } + {\mathbb{F}}^{ + } {\mathcal{L}}\left( {{\mathbb{F}}^{ - } } \right); $$
(A.19a)
$$ {\mathcal{L}}\left( {{\mathbb{F}}^{ \pm } } \right) = - \left[ {\frac{1}{2}\left\{ {{\mathcal{U}{\mathbb{P}}}_{lm} {\mathbf{B}}_{l} + {\omega {\mathbb{P}}}_{lm} {\mathbf{\mathcal{Z}}}_{l} } \right\}{\mathbf{A}}^{ \pm } {\mathbb{P}}_{lm}^{\text{T}} {\mathbb{C}}} \right]{\mathbb{M}}^{ - 1} . $$
(A.19b)

Linearization treatments are different for the full Stokes 4-vector case, and the scalar and Stokes 3-vector situations.

4-vector case. We apply linearization to both the left and right eigensystems:

$$ {\mathcal{L}}\left( {{\mathbb{X}}_{\alpha }^{*} } \right){\mathbb{G}} + {\mathbb{X}}_{\alpha }^{*} {\mathcal{L}}\left( {\mathbb{G}} \right) = 2k_{\alpha } {\mathcal{L}}\left( {k_{\alpha } } \right){\mathbb{X}}_{\alpha }^{*} + k_{\alpha }^{2} {\mathcal{L}}\left( {{\mathbb{X}}_{\alpha }^{*} } \right); $$
(A.20a)
$$ {{\mathbb{G}}\mathcal{L}}\left( {{\mathbb{X}}_{\alpha } } \right) + {\mathcal{L}}\left( {\mathbb{G}} \right){\mathbb{X}}_{\alpha } = 2k_{\alpha } {\mathcal{L}}\left( {k_{\alpha } } \right){\mathbb{X}}_{\alpha } + k_{\alpha }^{2} {\mathcal{L}}\left( {{\mathbb{X}}_{\alpha } } \right). $$
(A.20b)

We form a dot product \( \otimes \) by pre-multiplying the second of these equations by the transpose vector \( {\mathbb{X}}_{\alpha }^{*} \):

$$ {\mathbb{X}}_{\alpha }^{*} \otimes {{\mathbb{G}}\mathcal{L}}\left( {{\mathbb{X}}_{\alpha } } \right) + {\mathbb{X}}_{\alpha }^{*} \otimes {\mathcal{L}}\left( {\mathbb{G}} \right){\mathbb{X}}_{\alpha } = 2k_{\alpha } {\mathcal{L}}\left( {k_{\alpha } } \right){\mathbb{X}}_{\alpha }^{*} \otimes {\mathbb{X}}_{\alpha } + k_{\alpha }^{2} {\mathbb{X}}_{\alpha }^{*} \otimes {\mathcal{L}}\left( {{\mathbb{X}}_{\alpha } } \right). $$
(A.21)

Using the relation \( {\mathbb{X}}_{\alpha }^{*} \otimes {{\mathbb{G}}\mathcal{L}}\left( {{\mathbb{X}}_{\alpha } } \right) = {\mathbb{X}}_{\alpha }^{*} {\mathbb{G}} \otimes {\mathcal{L}}\left( {{\mathbb{X}}_{\alpha } } \right) = k_{\alpha }^{2} {\mathbb{X}}_{\alpha }^{*} \otimes {\mathcal{L}}\left( {{\mathbb{X}}_{\alpha } } \right) \), we find that

$$ y_{\alpha } \equiv {\mathcal{L}}\left( {k_{\alpha } } \right) = \frac{{{\mathbb{X}}_{\alpha }^{*} \otimes {\mathcal{L}}\left( {\mathbb{G}} \right){\mathbb{X}}_{\alpha } }}{{2k_{\alpha } {\mathbb{X}}_{\alpha }^{*} \otimes {\mathbb{X}}_{\alpha } }}. $$
(A.22)

This is the linearization of the separation constants. Next, we substitute this result in Eq. (A.20a) to obtain the following linear algebra problem for each eigensolution linearization:

$$ {\mathbb{H}}_{\alpha } {\mathcal{L}}\left( {{\mathbb{X}}_{\alpha } } \right) = {\mathbb{C}}_{\alpha } ;\quad {\mathbb{H}}_{\alpha } = {\mathbb{G}} - k_{\alpha }^{2} {\mathbb{E}};\quad {\mathbb{C}}_{\alpha } = 2k_{\alpha } y_{\alpha } {\mathbb{X}}_{\alpha } - {\mathcal{L}}\left( {\mathbb{G}} \right){\mathbb{X}}_{\alpha } . $$
(A.23)

For real eigensolutions, this system has rank \( 4N_{d} \), and for complex solutions, the rank is \( 8N_{d} \).

Implementation of this system of equations “as is” is not possible due to the degeneracy of the eigenproblem, and we need additional constraints to find the unique solution for \( { \mathcal{L}}\left( {{\mathbb{X}}_{\alpha } } \right) \). The treatment for real and complex solutions is different.

  • For the real-valued eigensolutions, the unit-modulus eigenvector normalization is \( {\mathbb{X}}_{\alpha } \otimes {\mathbb{X}}_{\alpha } = 1 \) in dot-product notation. Linearizing, this yields one equation:

$$ {\mathcal{L}}\left( {{\mathbb{X}}_{\alpha } } \right) \otimes {\mathbb{X}}_{\alpha } + {\mathbb{X}}_{\alpha } \otimes {\mathcal{L}}\left( {{\mathbb{X}}_{\alpha } } \right) = 0. $$
(A.24)

The solution procedure uses \( 4N_{d} - 1 \) equations from Eq. (A.23), along with Eq. (A.24) to form a slightly modified linear system of rank \( 4N_{d} \). This system is then solved by standard means using the DGETRF and DGETRS LU-decomposition routines from the LAPACK suite.

  • For complex-valued eigensolutions, Eq. (A.23) is a complex-variable system for both the real and imaginary parts of the linearized eigenvectors. There are \( 8N_{d} \) equations in all, but now we require two constraint conditions to remove the eigenproblem arbitrariness. The first is Eq. (24). The second condition is imposed by the following normalization from the LAPACK DGEEV routine for solution of complex-valued eigenproblems: for that element of a (complex-valued) eigenvector which has the largest real value, the corresponding imaginary part is always set to zero. Thus, for eigenvector \( {\mathbb{X}}_{\alpha } \) with components \( {\text{X}}_{j} \in {\mathbb{X}}_{\alpha } , j = 1,2, \ldots 4N_{d} \), if \( {\text{X}}_{q} \) satisfies condition \( {\text{Re}}[{\text{X}}_{q} ] = \max_{{j = 1,2, \ldots 4N_{d} }} \left\{ {{\text{Re}}[{\text{X}}_{j} ]} \right\} \), then \( {\text{Im}}[{\text{X}}_{q} ] = 0 \), and consequently \( {\mathcal{L}}\left( {{\text{Im}}[{\text{X}}_{q} ]} \right) = 0 \). This is the second condition. The solution procedure is then: (1) in Eq. (A.24) to strike out the qth row and column in matrix \( {\mathbb{H}}_{\alpha } \) for which \( {\text{Im}}[{\text{X}}_{q} ] \) is zero, and the qth column in vector \( {\mathbb{C}}_{\alpha } \); and (2) in the resulting system of rank \( 8N_{d} - 1 \), replace one of the rows with the normalization constraint Eq. (A.24). \( {\mathcal{L}}\left( {{\mathbb{X}}_{\alpha } } \right) \) is then the solution of the resulting linear system.

  • Scalar and 3-vector case. Here the (real-valued) eigensolutions are obtained using eigensolver ASYMTX—this has no adjoint solution, so there is no determination of \( {\mathcal{L}}\left( {k_{\alpha } } \right) \) as in Eq. (A.22). Instead, we solve for variables \( \left\{ {{\mathcal{L}}\left( {k_{\alpha } } \right),{\mathcal{L}}\left( {{\mathbb{X}}_{\alpha } } \right)} \right\} \) using \( {\mathbb{H}}_{\alpha } {\mathcal{L}}\left( {{\mathbb{X}}_{\alpha } } \right) = 2k_{\alpha } {\mathcal{L}}\left( {k_{\alpha } } \right){\mathbb{X}}_{\alpha } - {\mathcal{L}}\left( {\mathbb{G}} \right){\mathbb{X}}_{\alpha } \) from above, plus the normalization condition to form a joint system of rank \( 3N_{d} + 1 \) (vector) or rank \( N_{d} + 1 \) (scalar).

Having derived the linearizations \( \left\{ {{\mathcal{L}}\left( {k_{\alpha } } \right),{\mathcal{L}}\left( {{\mathbb{X}}_{\alpha } } \right)} \right\} \), we complete this section by differentiating the auxiliary result in Eq. (A.16) to establish \( { \mathcal{L}}\left( {{\mathbb{W}}_{\alpha }^{ \pm } } \right) \):

$$ {\mathcal{L}}\left( {{\mathbb{W}}_{\alpha }^{ \pm } } \right) = \frac{1}{2}{\mathbb{M}}^{ - 1} \left[ { \mp \frac{{{\mathcal{L}}\left( {k_{\alpha } } \right)}}{{k_{\alpha }^{2} }}{\mathbb{F}}^{ + } \pm \frac{1}{{k_{\alpha } }}{\mathcal{L}}\left( {{\mathbb{F}}^{ + } } \right)} \right]{\mathbb{X}}_{\alpha } + \frac{1}{2}{\mathbb{M}}^{ - 1} \left[ {{\mathbb{E}} \pm \frac{1}{{k_{\alpha } }}{\mathbb{F}}^{ + } } \right]{\mathcal{L}}\left( {{\mathbb{X}}_{\alpha } } \right). $$
(A.25)

Finally, we have linearizations of the transmittance derivatives in Eq. (A.17a):

$$ {\mathcal{L}}\left( {\exp \left[ { - k_{\alpha } x} \right]} \right) = - \left[ {x{\mathcal{L}}\left( {k_{\alpha } } \right) + k_{\alpha } {\mathcal{L}}\left( x \right)} \right]\exp \left[ { - k_{\alpha } x} \right]. $$
(A.26)

Since the partial layer optical thickness \( x \) is proportional to the total layer optical depth \( \Delta \) in an optically uniform layer, we have \( {\mathcal{L}}\left( x \right) = x/\Delta {\mathcal{L}}\left( \Delta \right) = x/\Delta {\mathcal{V}} \) in terms of the basic linearized optical property input \( {\mathcal{V}} \equiv \xi \partial \Delta /\partial \xi \).

1.3 A.3 Particular Integral Solar Solutions in VLIDORT

In VLIDORT, particular integrals for both solar and thermal sources in the vector RTE are established using traditional substitution methods, rather than the Green’s function approach which is used in LIDORT. This is mainly for bookkeeping reasons associated with the use of complex and real variables. In this section we discuss solutions for the solar sources using this method.

1.3.1 A.3.1 Solar-Beam SourcesSubstitution Method

For the solar source term with direction \( \left\{ { - \mu_{0} ,\phi_{0} } \right\} \), we have from Eq. (A.11) above, the following source terms in the discrete ordinate directions:

$$ {\mathbf{Q}}_{nm}^{ \odot } \left( {x, \pm \mu_{i} } \right) = \frac{{\omega_{n} }}{2}\sum\limits_{l = m}^{M} {{\mathbf{P}}_{l}^{m} \left( { \pm \mu_{i} } \right){\mathbf{B}}_{nl} {\mathbf{P}}_{l}^{m} \left( { - \mu_{0} } \right){\mathbf{F}}_{ \odot } {\text{T}}_{n} { \exp }\left[ { - \lambda_{n} x} \right]} . $$
(A.27)

Here, we have kept the layer index explicit, and \( x \) denotes the vertical optical thickness as measured from the top of layer \( n \); we used the pseudo-spherical treatment of solar beam attenuation (Sect. 3.4). The exponential form for the beam attenuation allows us to write the particular solution in the form:

$$ {\mathbf{I}}_{nm}^{ \odot } \left( {x, \pm \mu_{i} } \right) = {\mathbf{Z}}_{n} \left( { \pm \mu_{i} } \right){\text{T}}_{n} { \exp }\left[ { - \lambda_{n} x} \right], $$
(A.28)

and by analogy with the homogeneous case, we may define the following vectors of rank \( 4N_{d} \):

$$ {\mathbb{Z}}_{n}^{ \pm } = \left[ {{\mathbf{Z}}_{n}^{\text{T}} \left( { \pm \mu_{1} } \right),{\mathbf{Z}}_{n}^{\text{T}} \left( { \pm \mu_{2} } \right), \ldots ,{\mathbf{Z}}_{n}^{\text{T}} \left( { \pm \mu_{{N_{d} }} } \right)} \right]^{\text{T}} . $$
(A.29)

We decouple the equations using sum and difference vectors \( {\mathbb{R}}_{n} = {\mathbb{Z}}_{n}^{ + } + {\mathbb{Z}}_{n}^{ - } \) and \( {\mathbb{S}}_{n} = {\mathbb{Z}}_{n}^{ + } - {\mathbb{Z}}_{n}^{ - } \), and the order is reduced from \( 8N_{d} \) to \( 4N_{d} \). We obtain the following linear-algebra problem of rank \( 4N_{d} \):

$$ {\mathbb{H}}_{{\varvec{ n}}} {\mathbb{R}}_{n} = {\mathbb{B}}_{\varvec{n}} ; \quad {\mathbb{H}}_{\varvec{n}} = \lambda_{n}^{2} {\mathbb{E}} - {\mathbb{G}}_{n} ; \quad {\mathbb{B}}_{\varvec{n}} = \left[ {{\mathbb{F}}_{n}^{ - } {\mathbb{Q}}_{n}^{ + } + \frac{1}{{\lambda_{n} }}{\mathbb{Q}}_{n}^{ - } } \right]{\mathbb{M}}^{ - 1} ; $$
(A.30)
$$ {\mathbb{Q}}_{n}^{ \pm } = \omega_{n} \sum\nolimits_{l = m}^{M} {{\mathbb{P}}_{ \odot } \left( {l,m} \right){\mathbf{B}}_{nl} {\mathbf{A}}^{ \pm } {\mathbb{P}}_{ \odot }^{\text{T}} \left( {l,m} \right){\mathbb{M}}^{ - 1} ;} $$
(A.31a)
$$ {\mathbb{P}}_{ \odot } \left( {l,m} \right) = Diag\left[ {{\mathbf{P}}_{l}^{m} \left( { - \mu_{0} } \right),{\mathbf{P}}_{l}^{m} \left( { - \mu_{0} } \right), \ldots ,{\mathbf{P}}_{l}^{m} \left( { - \mu_{0} } \right)} \right]^{\text{T}} . $$
(A.31b)

Matrices \( {\mathbb{F}}_{n}^{ - } \) and \( {\mathbb{G}}_{n} \) were defined in Eqs. (A.15).

This system is solved using the LU-decomposition modules DGETRF and DGETRS from LAPACK; the formal solution is \( {\mathbb{X}}_{n}^{ \odot } = {\mathbb{H}}_{n}^{ - 1} {\mathbb{B}}_{\varvec{n}} \). The particular integral is completed with the auxiliary equations:

$$ {\mathbb{Z}}_{n}^{ \pm } = \frac{1}{2}{\mathbb{M}}^{ - 1} \left[ {{\mathbb{E}} \pm \frac{1}{{\lambda_{n} }}{\mathbb{F}}_{n}^{ + } } \right]{\mathbb{R}}_{n} . $$
(A.32)

In the scalar LIDORT model, this system has rank \( N_{d} \). In the vector model, the particular solution consists only of real variables.

1.3.2 A.3.2 Linearized Solar-Beam SourcesSubstitution Method

For this linearization, the most important consideration is the presence of cross-derivatives: in a fully illuminated atmosphere, the particular solution in layer \( n \) is differentiable with respect to atmospheric variables \( \xi_{p} \) in all layers \( p \le n \). For the solar beam attenuation, transmittance \( {\text{T}}_{n} \) depends on variables \( \xi_{p} \) in layers \( p < n \), while average secant \( \lambda_{n} \) depends on variables \( \xi_{p} \) in layers \( p \le n \). Linearization of the average-secant parameterization is in Sect. 3.4. It follows that the solution vectors \( {\mathbb{Z}}_{n}^{ \pm } \) will also depend on \( \xi_{p} \) for \( p \le n \), so their linearizations will contain cross-derivatives. Finally, we note that the eigenmatrix \( {\mathbb{G}}_{n} \) is constructed from optical properties only defined in layer \( n \), so that \( {\mathcal{L}}_{p} \left( {{\mathbb{G}}_{n} } \right) = 0, \forall p \ne n \).

Differentiation of Eqs. (A.30) and A.31a) yields a related linear problem:

$$ {\mathbb{H}}_{{\varvec{ n}}} {\mathcal{L}}_{p} \left( {{\mathbb{R}}_{n} } \right) = {\mathbb{B}}_{np}^{{\prime }} = {\mathcal{L}}_{p} \left( {{\mathbb{B}}_{\varvec{n}} } \right) - {\mathcal{L}}_{p} \left( {{\mathbb{H}}_{{\varvec{ n}}} } \right){\mathbb{R}}_{n} ; $$
(A.33a)
$$ {\mathcal{L}}_{p} \left( {{\mathbb{H}}_{{\varvec{ n}}} } \right) = - \delta_{np} {\mathcal{L}}_{p} \left( {{\mathbb{G}}_{n} } \right) + 2\lambda_{n} {\mathcal{L}}_{p} \left( {\lambda_{n} } \right){\mathbb{E}}; $$
(A.33b)
$$ {\mathcal{L}}_{p} \left( {{\mathbb{B}}_{\varvec{n}} } \right) = \delta_{np} \left[ {{\mathcal{L}}_{n} \left( {{\mathbb{F}}_{n}^{ - } } \right){\mathbb{Q}}_{n}^{ + } + {\mathbb{F}}_{n}^{ - } {\mathcal{L}}_{n} \left( {{\mathbb{Q}}_{n}^{ + } } \right) + \frac{1}{{\lambda_{n} }}{\mathcal{L}}_{n} \left( {{\mathbb{Q}}_{n}^{ - } } \right)} \right]{\mathbb{M}}^{ - 1} - \frac{{{\mathcal{L}}_{p} \left( {\lambda_{n} } \right)}}{{\lambda_{n}^{2} }}{\mathbb{Q}}_{n}^{ - } {\mathbb{M}}^{ - 1} ; $$
(A.33c)
$$ {\mathcal{L}}_{n} \left( {{\mathbb{Q}}_{n}^{ \pm } } \right) = \sum\nolimits_{l = m}^{M} {\left[ {{\mathcal{U}}_{n} {\mathbb{P}}_{ \odot } \left( {l,m} \right){\mathbf{B}}_{nl} + \omega_{n} {\mathbb{P}}_{ \odot } \left( {l,m} \right){\boldsymbol{\mathcal{Z}}}_{nl} } \right]} {\mathbf{A}}^{ \pm } {\mathbb{P}}_{ \odot }^{\text{T}} \left( {l,m} \right){\mathbb{M}}^{ - 1} . $$
(A.33d)

In Eq. (A.33c), the quantity \( {\mathcal{L}}_{n} \left( {{\mathbb{F}}_{n}^{ - } } \right) \) comes from Eq. (A.19b). This linear system has the same matrix \( {\mathbb{H}}_{{\varvec{ }n}} \), but with a different source vector \( {\mathbb{B}}_{np}^{{\prime }} \). The solution is then found by back-substitution, given that the inverse of \( {\mathbb{H}}_{{\varvec{ }n}} \) has already been established when solving for \( {\mathbb{R}}_{n} \). Thus, \( {\mathcal{L}}_{p} \left( {{\mathbb{R}}_{n} } \right) = {\mathbb{H}}_{n}^{ - 1} {\mathbb{B}}_{n}^{{\prime }} \). Linearization of the particular integral is then completed through differentiation of the auxiliary equations (A.32):

$$ {\mathcal{L}}_{p} \left( {{\mathbb{Z}}_{n}^{ \pm } } \right) = \frac{1}{2}{\mathbb{M}}^{ - 1} \left[ {{\mathbb{E}} \pm \frac{1}{{\lambda_{n} }}{\mathbb{F}}_{n}^{ + } } \right]{\mathcal{L}}_{p} \left( {{\mathbb{R}}_{n} } \right) \pm \frac{1}{{2\lambda_{n}^{2} }}{\mathbb{M}}^{ - 1} \left[ {\lambda_{n} \delta_{np} {\mathcal{L}}_{n} \left( {{\mathbb{F}}_{n}^{ + } } \right) - {\mathcal{L}}_{p} \left( {\lambda_{n} } \right){\mathbb{F}}_{n}^{ + } } \right]{\mathbb{R}}_{n} . $$
(A.34)

Appendix B. BRDF Kernel Functions

2.1 B.1 Ocean Glitter Kernels

2.1.1 B.1.1 Cox-Munk Glint Reflectance

For ocean glitter, we use the well-known geometric-optics regime for a single rough-surface redistribution of incident light, in which the reflection function is governed by Fresnel reflectance and takes the form (Jin et al. 2006):

$$ \rho_{CM} \left( {\mu ,\mu^{{\prime }} ,\varphi - \varphi^{{\prime }} ,m,\sigma^{2} } \right) = \frac{{F\left( {m,\theta_{r} } \right)}}{{\mu \mu^{{\prime }} \left| {\gamma_{r} } \right|^{4} }}P\left( {\gamma_{r} ,\sigma^{2} } \right)D\left( {\mu ,\mu^{{\prime }} ,\sigma^{2} } \right) $$
(B.1)

Here, \( \sigma^{2} \) is the slope-squared variance (also known as the mean-slope-square) of the Gaussian probability function \( P\left( {\gamma_{r} ,\sigma^{2} } \right) \) which has argument \( \gamma_{r} \) (the polar direction of the reflected beam), \( D\left( {\mu ,\mu^{{\prime }} ,\sigma^{2} } \right) \) is a shadow function correction (Sancer 1969).

\( F\left( {m,\theta_{r} } \right) \) is the scalar Fresnel reflection for incident angle \( \theta_{r} = \frac{1}{2}\theta_{SCAT} \) and relative refractive index \( m \). The scattering angle \( \theta_{SCAT} \) is determined in the usual manner from the incident and reflected directional cosines \( \mu^{{\prime }} \) and \( \mu \), and the relative azimuth \( \varphi - \varphi^{{\prime }} \).

The two non-linear parameters characterizing this kernel are \( \left\{ {m,\sigma^{2} } \right\} \). The probability and shadow functions are given by:

$$ \begin{aligned} P\left( {\alpha ,\sigma^{2} } \right) & = \frac{1}{{\pi \sigma^{2} }}{ \exp } \left[ { - \frac{{\alpha^{2} }}{{\sigma^{2} \left( {1 - \alpha^{2} } \right)}}} \right].\quad D\left( {\alpha ,\beta ,\sigma^{2} } \right) = \frac{1}{{1 +\Lambda \left( {\alpha ,\sigma^{2} } \right) +\Lambda \left( {\beta ,\sigma^{2} } \right)}}; \\\Lambda \left( {\alpha ,\sigma^{2} } \right) & = \frac{1}{2}\left\{ {\sqrt {\frac{{\left( {1 - \alpha^{2} } \right)}}{\pi }} \frac{\sigma }{\alpha }{ \exp } \left[ { - \frac{{\alpha^{2} }}{{\sigma^{2} \left( {1 - \alpha^{2} } \right)}}} \right] - erfc\left[ { - \frac{\alpha }{{\sigma \sqrt {\left( {1 - \alpha^{2} } \right)} }}} \right]} \right\}. \\ \end{aligned} $$
(B.2)

The variance is commonly related to the wind speed \( W \) in (m/s) through the empirical relation \( \sigma^{2} = 0.003 + 0.00512W \) (Cox and Munk 1954a). A typical value for \( m \) is 1.33.

For the linearization, the key parameter is the wind-speed (or equivalently, the mean slope-square) The probability function is easily differentiated with respect to \( \sigma^{2} \). Indeed, we have:

$$ \frac{{\partial P\left( {\alpha ,\sigma^{2} } \right)}}{{\partial \sigma^{2} }} = \frac{{P\left( {\alpha ,\sigma^{2} } \right)}}{{\sigma^{4} }}\left[ {\frac{{\alpha^{2} }}{{\left( {1 - \alpha^{2} } \right)}} - \sigma^{2} } \right]. $$
(B.3)

Again, the shadow function can be differentiated analytically with respect to \( \sigma^{2} \) in a straightforward manner, once we recall that the derivative of the error function is Gaussian. We thus have linearized BRDF quantities prepared for (V)LIDORT, which will then be able to generate analytic weighting functions with respect to the wind speed.

Linearization of the kernel with respect to refractive index m will require the partial derivative \( \partial F\left( {m,\theta_{r} } \right)/\partial m \), which is easy to determine from the usual Fresnel formula; this Jacobian is less useful.

We note also that VLIDORT has a vector kernel function for sea-surface glitter reflectance, based on the specification in Mishchenko and Travis (1998).

This formulation is for a single Fresnel reflectance by wave facets. In reality, glitter is the result of many reflectances. It is possible to incorporate a correction for multiple reflectances for this glitter contribution, both for the direct-bounce term and the Fourier components. Given that the glitter maximum is typically dominated by direct reflectance of the solar beam, we confine discussion to multiple-reflectance of this term. We consider only one extra order of reflectance:

$$ R\left( {\Omega ,\Omega _{0} } \right) \cong R_{0} \left( {\Omega ,\Omega _{0} } \right) + R_{1} \left( {\Omega ,\Omega _{0} } \right); R_{1} \left( {\Omega ,\Omega _{0} } \right) = \int\limits_{0}^{2\pi } {\int\limits_{0}^{1} {R_{0} \left( {\Omega ,\Omega ^{{\prime }} } \right)R_{0} \left( {\Omega ^{{\prime }} ,\Omega _{0} } \right)d\mu^{{\prime }} d\varphi^{{\prime }} } } . $$
(B.4)

\( R_{0} \left( {\Omega ,\Omega ^{{\prime }} } \right) \) is the zeroth-order reflectance from incident direction \( \Omega ^{{\prime }} = \left\{ {\mu^{{\prime }} ,\varphi^{{\prime }} } \right\} \) to reflected direction \( \Omega = \left\{ {\mu ,\varphi } \right\} \). The azimuthal integration is done by double Gaussian quadrature over the intervals \( \left[ { - \pi ,0} \right] \) and \( \left[ {0,\pi } \right] \); the polar stream integration is also done by quadrature. It is obviously possible to calculate higher orders of reflectance for all BRDF functions, but this rapidly becomes computationally prohibitive. We have found that the neglect of multiple glitter reflectances can lead to errors of 1–3% in the upwelling intensity at the top of the atmosphere, the higher figures being for larger solar zenith angles. Finally, we note that the higher-order reflectances are in turn differentiable with respect to the slope-squared parameter, so that Jacobians for the wind speed can still be determined.

2.1.2 B.1.2 Alternative Cox-Munk Glint Reflectance

The above ocean-glint reflectance option in LIDORT is based on an older implementation of the well-known Cox-Munk distribution of wave facets—this does not include any treatment of whitecaps (foam), and there is no allowance for the wind direction. Further, the CM implementation is based on a fixed real-valued refractive index for water.

We have now added an alternative Cox-Munk implementation which addresses these issues. This new-CM option is based on the glint treatment in the 6S code (Vermote et al. 1997; Kotchenova et al. 2006), and includes an empirical whitecap contribution. The latter model also includes a more recent treatment of water-leaving radiance (Morel and Gentili 2009), and the LIDORT SLEAVE supplement has been updated according to the 6S treatment. These new 6S-based options in the LIDORT BRDF and SLEAVE supplements are designed to operate in tandem. Indeed, the total surface radiance in the 6S model is given by

$$ I\left( {\mu_{1} ,\mu_{0} ,\varphi_{1} - \varphi_{0} } \right) = \left( {1 - R_{wc} } \right)S\left( {\mu_{1} ,\mu_{0} } \right) + R_{wc} + \left( {1 - R_{L} } \right) \rho_{NCM} \left( {\mu_{1} ,\mu_{0} ,\varphi_{1} - \varphi_{0} } \right), $$
(B.5)

where the water-leaving term \( S\left( {\mu_{1} ,\mu_{0} } \right) \) does depend on the outgoing and incoming directions but is azimuth-independent, \( \rho_{NCM} \) is the Cox-Munk glint reflectance, and \( R_{wc} \) and \( R_{L} \) are the empirically-derived whitecap contributions (final and Lambertian).

For LIDORT to possess this functionality for the ocean surface, the BRDF supplement must provide the second and third terms on the right-hand-side of (B.5), while the SLEAVE supplement supplies the first term. Note that both supplements require the same whitecap formulation. The derivation of the water-leaving term is done in the next section.

The Cox-Munk calculation for \( \rho_{NCM} \) depends on the wind speed, wind-direction and refractive index. The latter is a complex variable that depends on the salinity of the ocean, and we use the 6S formulation. The anisotropic treatment assumes an (azimuthal) wind direction relative to the solar incident beam, and this follows the formulae developed by Cox and Munk in the 1950s.

For facet anisotropy, the wind-direction is dependent on the solar position, and it follows that any BRDF quantity will pick up additional dependence on the solar angle. It is then only possible to use this “New-CM” option with a single solar zenith angle (and hence a single wind-direction azimuth) - multiple solar beams are ruled out. See also the remark in Sect. 4.2.3 regarding use of the black-sky albedo. There is exception handling for this eventuality.

In addition, we have linearized this “New-CM” glint reflectance with respect to the wind speed. Finally, we note that this glint reflectance is scalar only, so that the above considerations apply only to the (1,1) element of the reflectance matrix in the vector BRDF supplement for VLIDORT. Polarization of this contribution is currently undergoing investigation.

2.2 B.2 Land-Surface BRDF Kernels

2.2.1 B.2.1 MODIS BRDF System

The five MODIS-type kernels (numbers 2–6 in Table 3) (Wanner et al. 1995) must be used in a linear combination with a Lambertian kernel. The kernels divide naturally into two groups: the volume scattering terms with no non-linear parameters (Ross-thin, Ross-thick) and the geometric-optics terms with 2 non-linear parameters (Li-sparse, Li-dense) or no non-linear parameters (Roujean). See Wanner et al. (1995) and Spurr (2004) for details of the kernel formulae.

In fact, it is standard practice in MODIS BRDF retrievals to use a combination of Lambertian, Ross-thick and Li-Sparse kernels, and this 3-kernel combination is common:

$$ \rho_{total} \left( {\Omega ,\Omega ^{{\prime }} } \right) = f_{iso} + f_{vol} \rho_{RossThick} \left( {\Omega ,\Omega ^{{\prime }} } \right) + f_{geo} \rho_{LiSparse} \left( {\Omega ,\Omega ^{{\prime }} } \right) $$
(B.6)

An alternative form of the Ross kernels has also been introduced—this has a better parameterization of the hot-spot effect. The Rahman and Hapke kernels (#7 or #8) were discussed in [R1].

2.2.2 B.2.2 BPDF Kernels

The 3 BPDF kernels (Maignan et al. 2009) (numbers 10–12 in Table 3) were developed as semi-empirical models for polarized land-surface bidirectional reflectances. All three kernels are based on Fresnel reflectance. The polarization effects enter through the Fresnel reflectance term; for LIDORT, we require only the scalar Fresnel reflectance.

For the BPDF “Soil” type, the scalar reflectance is:

$$ \rho_{SOIL} \left( {\mu ,\mu^{{\prime }} ,\varphi - \varphi^{{\prime }} ,m} \right) = \frac{{F\left( {m,\theta_{r} } \right)}}{{4\mu \mu^{{\prime }} }}\left( {1 - \sin \theta_{r} } \right). $$
(B.7)

As before, \( F\left( {m,\theta_{r} } \right) \) is the Fresnel reflection for incident angle \( \theta_{r} = \frac{1}{2}\theta_{SCAT} \) and relative refractive index \( m \) (which is the sole kernel parameter). Linearization of the kernel with respect to refractive index m requires the partial derivative of \( F\left( {m,\theta_{r} } \right) \).

For the BDPF “Vegetation” type, there is dependence on the leaf orientation probability \( \sigma \left( \alpha \right) \) and the leaf facet projections, through use of a plagiophile distribution:

$$ \begin{aligned} \rho_{VEGN} \left( {\mu ,\mu^{{\prime }} ,\varphi - \varphi^{{\prime }} ,m} \right) & = \frac{{F\left( {m,\theta_{r} } \right)}}{{4\mu \mu^{{\prime }} }}\frac{\zeta \left( \alpha \right)}{\text{H}}\left( {1 - \sin \theta_{r} } \right); \\ \cos \alpha & = \frac{{\left( {\mu + \mu^{{\prime }} } \right)}}{{2\cos \theta_{r} }}; \quad \zeta \left( \alpha \right) = \frac{16}{\pi } \cos^{2} \alpha \,\sin \alpha ; \\ {\text{H}} & = \frac{{\mathop \sum \nolimits_{k = 0}^{3} a_{k} \mu^{k} }}{\mu } + \frac{{\mathop \sum \nolimits_{k = 0}^{3} a_{k} \mu^{{{\prime }k}} }}{{\mu^{{\prime }} }}, \\ \end{aligned} $$
(B.8)

where the leaf projection \( {\text{H}} \) depends on the set of “plagiophile coefficients” \( \left\{ {a_{k} } \right\} \). Again, the refractive index is the only surviving kernel parameter to be considered for linearization.

For the BPDF “NDVI” kernel, we have:

$$ \rho_{NDVI} \left( {\mu ,\mu^{{\prime }} ,\varphi - \varphi^{{\prime }} ,m,N,C} \right) = \frac{{CF\left( {m,\theta_{r} } \right)}}{{4\left( {\mu + \mu^{\prime}} \right)}}\exp \left[ { - \tan \theta_{r} } \right]\exp \left[ { - N} \right], $$
(B.9)

where there are exponential attenuation terms, one of which depends on the NDVI \( N \); in this formula, the scaling factor is \( C \) (nominally, this is set to 1.0). Linearizations with respect to the parameters \( N \) and \( C \) are easy to establish. The NDVI varies from −1 to 1 and is defined as the ratio of the difference to the sum of two radiance measurements, one in the visible and one in the infrared.

Appendix C. Taylor Series Expansions

We have already noted that certain “multipliers” arising from optical-thickness integrations which are needed to find various solutions to components of the RT field may possess instability when certain limiting conditions are in place. Taylor series expansions are then required to remove such instabilities. Although some simple expansions were used in earlier versions of the LIDORT and VLIDORT codes, the whole issue has been completely revised in recent versions of the code (since [R1] was published), and we go into some detail here. Remarks on the VLIDORT implementation are made where appropriate.

3.1 C.1 Multipliers for the Intensity Field

We first look at the homogeneous-solution and primary-scatter downwelling multipliers, and the Green’s function (downwelling) multiplier for the diffuse source term at discrete-ordinate polar cosines. These are respectively:

$$ H_{j}^{ \downarrow } = \frac{{\left( {e^{{ - \Delta k_{j} }} - e^{{ - \Delta \mu^{ - 1} }} } \right)}}{{\mu^{ - 1} - k_{j} }}; F^{ \downarrow } = \frac{{\left( {e^{{ - \Delta \mu^{ - 1} }} - e^{ - \Delta \lambda } } \right)}}{{\lambda - \mu^{ - 1} }}; G_{j}^{ \downarrow } = \frac{{\left( {e^{{ - \Delta k_{j} }} - e^{ - \Delta \lambda } } \right)}}{{\lambda - k_{j} }}. $$
(C.1)

As before, \( \lambda \) is the layer average secant corresponding to spherical-shell attenuation of the solar beam, \( k_{j} \) is the separation constant corresponding to discrete ordinate stream \( j \) arising from solution of the homogeneous RTE eigenproblem, \( \mu \) is the cosine of the line-of-sight viewing polar angle, and \( \Delta \) is the layer total optical depth. [Layer index \( n \) is suppressed for clarity].

For polarized RT with VLIDORT, some of the separation constants \( k_{j} \) may be complex variables; but \( \lambda \) and \( \mu \) are always real-valued. Taylor series expansions involving \( k_{j} \) are only applicable for real values.

We consider the limiting cases \( \left| {\lambda - k_{j} } \right| \to 0, \left| {\mu^{ - 1} - k_{j} } \right| \to 0 \) or \( \left| {\lambda - \mu^{ - 1} } \right| \to 0 \). Writing \( \epsilon \) for any of these quantities, if the order of the Taylor-series expansion is \( M \), then we neglect terms \( O\left( {\epsilon^{M + 1} } \right) \). Considering first the multiplier \( G_{j}^{ \downarrow } \) from Eq. (C.1), then if \( \epsilon = \lambda - k_{j} \), we have \( e^{{ - \Delta k_{j} }} = We^{\epsilon \Delta } \approx W{\mathbf{z}}_{M + 1} \left( \Delta \right) \), where we have written \( W = e^{ - \Delta \lambda } \), and we have used the notation \( {\mathbf{z}}_{M + 1} \left( x \right) = \sum\nolimits_{m = 0}^{M + 1} {z_{m} \left( x \right)\epsilon^{m} } \) to approximate the exponential \( e^{\epsilon \Delta } \), so that the coefficients are \( z_{0} \left( x \right) = 1, z_{m} \left( x \right) = x^{m} /m! \) for \( 0 < m \le M + 1 \). Applying the expansion, we find:

$$ G_{j}^{ \downarrow } \approx \frac{{W\left( {{\mathbf{z}}_{M + 1} \left( \Delta \right) - 1} \right)}}{\epsilon } = \Delta W{\mathbf{z}}_{M}^{*} \left( \Delta \right). $$
(C.2)

Here, the new coefficients are \( z_{0}^{*} = 1, z_{m}^{*} = \Delta^{m} /\left( {m + 1} \right)! \) for \( 0 < m \le M \). Note that we need to expand first to order \( M + 1 \) to ensure that the final expression is defined to order \( M \).

Next we look at those two multipliers for the Green’s function post-processed field which may be susceptible to instability through appearance of the term \( \lambda - k_{j} \) in the denominator:

$$ M_{j}^{ \uparrow \uparrow } = \frac{{H_{j}^{ \uparrow } - F^{ \uparrow } }}{{\lambda - k_{j} }};\quad M_{j}^{ \downarrow \downarrow } = \frac{{H_{j}^{ \downarrow } - F^{ \downarrow } }}{{\lambda - k_{j} }}, $$
(C.3)

where \( H_{j}^{ \downarrow } \) and \( F^{ \downarrow } \) have been defined in Eq. (1), and

$$ H_{j}^{ \uparrow } = \frac{{\left( {1 - e^{{ - \Delta k_{j} }} e^{{ - \Delta \mu^{ - 1} }} } \right)}}{{\mu^{ - 1} + k_{j} }}; F^{ \uparrow } = \frac{{\left( {1 - e^{{ - \Delta \mu^{ - 1} }} e^{ - \Delta \lambda } } \right)}}{{\lambda + \mu^{ - 1} }}. $$
(C.4)

Looking at \( M_{j}^{ \uparrow \uparrow } \) in Eq. (C.3), we set \( k_{j} = \lambda - \epsilon \), and find:

$$ M_{j}^{ \uparrow \uparrow } \approx \frac{Y}{\epsilon }\left[ {\left( {1 - WU{\mathbf{z}}_{M + 1} \left( \Delta \right)} \right){\mathbf{a}}_{M + 1} \left( Y \right) - \left( {1 - WU} \right)} \right], $$
(C.5)

where \( U = e^{{ - \Delta \mu^{ - 1} }} ,Y = \left( {\lambda + \mu^{ - 1} } \right)^{ - 1} \), and series \( {\mathbf{a}}_{M + 1} \left( x \right) \) has coefficients \( a_{0} \left( x \right) = 1, a_{m} \left( x \right) = x^{m} \) for \( 0 < m \le M + 1 \). Since \( a_{0} = z_{0} = 1 \), it is apparent that the \( \left( {1 - WU} \right) \) contributions will fall out, and the result can then be written:

$$ M_{j}^{ \uparrow \uparrow } \approx Y\left[ {Y{\mathbf{a}}_{M} \left( Y \right) - WU{\mathbf{c}}_{M} \left( {\Delta ,Y} \right)} \right]. $$
(C.6)

Here \( {\mathbf{c}}_{M} \left( {\Delta ,Y} \right) \) has coefficients \( c_{0} = 1 \), and \( c_{m} \left( {\Delta ,Y} \right) = \sum\nolimits_{p = 0}^{m} {z_{p} \left( \Delta \right)a_{m - p} \left( Y \right)} \) for \( 0 < m \le M \). We have found that use of these series coefficients is convenient for computation, allowing us to generate expressions to any order of accuracy without complicated algebraic expressions.

The other multiplier \( M_{j}^{ \downarrow \downarrow } \) in Eq. (C.3) may be treated similarly. Note that multipliers in Eq. (C.4) are numerically stable entities, along with the other Green’s function multipliers \( G_{j}^{ \uparrow } , M_{j}^{ \downarrow \uparrow } \) and \( M_{j}^{ \uparrow \downarrow } \), the latter three quantities being defined with denominator \( \lambda + k_{j} \).

The above multipliers are required for output of the upwelling and downwelling radiance fields at layer boundaries. LIDORT has the ability to generate output at any level away from layer boundaries (the “partial-layer” option). In this case, source function integration to an optical thickness \( \tau < \Delta \) in layer n will result in “partial-layer” multipliers similar to those already defined above. For example, consider the following three functions:

$$ H_{j}^{ \downarrow } \left( \tau \right) = \frac{{\left( {e^{{ - \tau k_{j} }} - e^{{ - \tau \mu^{ - 1} }} } \right)}}{{\mu^{ - 1} - k_{j} }}; F^{ \downarrow } \left( \tau \right) = \frac{{\left( {e^{{ - \tau \mu^{ - 1} }} - e^{ - \tau \lambda } } \right)}}{{\lambda - \mu^{ - 1} }}; G_{j}^{ \downarrow } \left( \tau \right) = \frac{{\left( {e^{{ - \tau k_{j} }} - e^{ - \tau \lambda } } \right)}}{{\lambda - k_{j} }}; $$
(C.7)

These are the downwelling partial-layer multipliers equivalent to those in Eq. (C.1). Taylor series expansions for these and the corresponding Green’s function multipliers have been generated in a similar fashion.

3.2 C.2 Linearized Multipliers for the Jacobian Fields

The entire LIDORT discrete ordinate solution is analytically differentiable (Spurr 2002) with respect to any atmospheric quantity, and this includes the multipliers discussed above. It is therefore necessary to develop Taylor series expansions for those linearized multipliers which are susceptible to numerical instability.

We start with partial derivatives \( \dot{k}_{j} \equiv \partial k_{j} /\partial \xi ,\dot{\lambda } \equiv \partial \lambda /\partial \xi \) and \( \dot{\Delta } \equiv \partial \Delta /\partial \xi \) with respect to some atmospheric quantity \( \xi \) defined in layer \( n \) (index is not explicit here). Since the layer optical depth \( \Delta \) is an intrinsic optical property, its derivative \( \dot{\Delta } \) is also an intrinsic input. The viewing angle cosine \( \mu \) has no derivative. Note that \( \dot{k}_{j} = 0 \) and \( \dot{\Delta } = 0 \) for a profile atmospheric quantity \( \xi_{m } \) defined in layer \( m \ne n \) (no cross-layer derivatives); on the other hand, the average secant \( \lambda \) will have cross-layer derivatives for layers \( m < n \), thanks to solar beam attenuation through the atmosphere. For the third multiplier in Eq. (C.1), the derivative is

$$ \frac{{\partial G_{j}^{ \downarrow } }}{\partial \xi } = \frac{{ - e^{{ - \Delta k_{j} }} \left( {\dot{k}_{j} \Delta + k_{j} \dot{\Delta }} \right) + e^{ - \Delta \lambda } \left( {\dot{\lambda }\Delta + \lambda \dot{\Delta }} \right) - G_{j}^{ \downarrow } \left( {\dot{\lambda } - \dot{k}_{j} } \right)}}{{\lambda - k_{j} }}. $$
(C.8)

Expanding Eq. (C.8) as a Taylor series with \( k_{j} = \lambda - \epsilon \), and using the series-coefficient notation developed in Sect. C.1, we find:

$$ \frac{{ \partial G_{j}^{ \downarrow } }}{\partial \xi } \approx \frac{{ - \left( {\dot{k}_{j} \Delta + \lambda \dot{\Delta } + \epsilon \dot{\Delta }} \right)W{\mathbf{z}}_{M + 1} + \left( {\dot{\lambda }\Delta + \lambda \dot{\Delta }} \right)W - \left( {\dot{\lambda } - \dot{k}_{j} } \right)\Delta W{\mathbf{z}}_{M + 1}^{*} }}{\epsilon }. $$
(C.9)

Again, \( W = e^{ - \Delta \lambda } \), and the series \( {\mathbf{z}}_{M + 1} \) and \( {\mathbf{z}}_{M + 1}^{*} \) have argument \( {\Delta } \) and were defined in Sect. C.1. Since \( z_{0} = z_{0}^{*} = 1 \), the lowest-order terms in the numerator of Eq. (C.9) will fall out, and we are left with:

$$ \frac{{\partial G_{j}^{ \downarrow } }}{\partial \xi } \approx - W\left[ {\dot{\Delta }{\mathbf{z}}_{M} + \left( {\dot{k}_{j} \Delta + \lambda \dot{\Delta }} \right)\Delta {\mathbf{z}}_{M}^{*} + \left( {\dot{\lambda } - \dot{k}_{j} } \right)\Delta^{2} {\mathbf{z}}_{M}^{\dag } } \right], $$
(C.10)

where the third series \( {\mathbf{z}}_{M}^{\dag } \) has coefficients \( z_{0}^{\dag } = 1/2 \) and \( z_{m}^{\dag } = \Delta^{m} /\left( {m + 2} \right)! \) for \( 0 < m \le M \). Note that the presence of the series \( {\mathbf{z}}_{M}^{\dag } \) to order \( M \) implies that the original series must be computed to order \( M + 2 \); that is, we require \( {\mathbf{z}}_{M + 2} \).

Linearization of the Green’s function multipliers in Eq. (C.3) follows similar considerations. We give one example; explicit differentiation of \( M_{j}^{ \uparrow \uparrow } \) yields

$$ \frac{{\partial M_{j}^{ \uparrow \uparrow } }}{\partial \xi } = \frac{{\dot{H}_{j}^{ \uparrow } - \dot{F}^{ \uparrow } - \left( {\dot{\lambda } - \dot{k}_{j} } \right)M_{j}^{ \uparrow \uparrow } }}{{\lambda - k_{j} }}; $$
(C.11)
$$ \dot{H}_{j}^{ \uparrow } = \frac{{Ue^{{ - \Delta k_{j} }} \left[ {\dot{\Delta }\left( {\mu^{ - 1} + k_{j} } \right) + \dot{k}_{j} \Delta } \right] - \dot{k}_{j} H_{j}^{ \uparrow } }}{{\mu^{ - 1} + k_{j} }}; \dot{F}^{ \uparrow } = \frac{{UW\left[ {\dot{\Delta }\left( {\mu^{ - 1} + \lambda } \right) + \dot{\lambda }\Delta } \right] - \dot{\lambda }F^{ \uparrow } }}{{\lambda + \mu^{ - 1} }}. $$
(C.12)

Expanding in the usual manner and employing results established already, we find

$$ \dot{H}_{j}^{ \uparrow } \approx \left[ {UW\left( {\dot{k}_{j} \Delta + \dot{\Delta }Y^{ - 1} - \epsilon \dot{\Delta }} \right){\mathbf{z}}_{N} - \dot{k}_{j} Y\left( {1 - WU{\mathbf{z}}_{N} } \right){\mathbf{a}}_{N} } \right]Y{\mathbf{a}}_{N} ; $$
(C.13)
$$ \dot{F}^{ \uparrow } = \left[ {UW\left( {\dot{\lambda }\Delta + \dot{\Delta }Y^{ - 1} } \right) - \dot{\lambda }Y\left( {1 - WU} \right)} \right]Y; $$
(C.14)

where now \( N = M + 2 \) terms in the series have been retained in the expansions. Eq. (C.11) implies that we also require the original multiplier \( M_{j}^{ \uparrow \uparrow } \) expanded to order \( M + 1 \), that is,

$$ M_{j}^{ \uparrow \uparrow } \approx Y\left[ {Y{\mathbf{a}}_{M + 1} \left( Y \right) - WU{\mathbf{c}}_{M + 1} \left( {\Delta ,Y} \right)} \right] $$
(C.15)

should be used in Eq. (C.11). Putting together the above three equations, and canceling out the lowest-order terms in the expansions, we find after some algebra that

$$ \frac{{\partial M_{j}^{ \uparrow \uparrow } }}{\partial \xi } \approx Y\left[ {WU{\mathbf{S}}_{M}^{\left( 1 \right)} - {\mathbf{S}}_{M}^{\left( 2 \right)} - {\mathbf{S}}_{M}^{\left( 3 \right)} } \right], $$
(C.16)

where the three series \( {\mathbf{S}}_{M}^{\left( q \right)} = \sum\nolimits_{m = 0}^{M} {s_{m}^{\left( q \right)} \epsilon^{m } } \) have coefficients

$$ s_{m}^{\left( 1 \right)} = \left( {\dot{\Delta }Y^{ - 1} + \dot{k}_{j} \Delta } \right)c_{m + 1} \left( {\Delta ,Y} \right) - \dot{\Delta }c_{m} \left( {\Delta ,Y} \right); $$
(C.17a)
$$ s_{m}^{\left( 2 \right)} = Y\dot{k}_{j} \left[ {b_{m + 1} \left( Y \right) + WUd_{m + 1} \left( {\Delta ,Y} \right)} \right]; $$
(C.17b)
$$ s_{m}^{\left( 3 \right)} = \left( {\dot{\lambda } - \dot{k}_{j} } \right)\left[ {b_{m + 2} \left( Y \right) + WUc_{m + 2} \left( {\Delta ,Y} \right)} \right]. $$
(C.17c)

Subsidiary coefficients for \( 0 < m \le M \) are given by \( a_{m} = Y^{m} \) (series expansion of \( \left( {1 - \epsilon Y} \right)^{ - 1} \)), and \( b_{m} = \left( {m + 1} \right)Y^{m} \) (series expansion of \( \left( {1 - \epsilon Y} \right)^{ - 2} \)). We also have the product coefficients \( c_{m} \left( {\Delta ,Y} \right) = \sum\nolimits_{p = 0}^{m} {z_{p} \left( \Delta \right)a_{m - p} \left( Y \right)} \) and \( d_{m} \left( {\Delta ,Y} \right) = \sum\nolimits_{p = 0}^{m} {z_{p} \left( \Delta \right)b_{m - p} \left( Y \right)} \) obtained from the first exponential series \( {\mathbf{z}}\left( \Delta \right) \) which approximates \( e^{\Delta \epsilon } \).

Derivatives of the other multipliers subject to possible instability may be obtained similarly, and there are also derivatives of the partial-layer multipliers to be considered.

Software for all these instability cases has been written for LIDORT and VLIDORT - in both models, the order \( M \) controls the accuracy of the Taylor series expansions. The other parameter controlling the use of these limiting-case calculations is the “small-number” value \( \epsilon \). LIDORT and VLIDORT use double-precision floating-point arithmetic. With this in mind, we have chosen \( \varepsilon = 10^{ - 3} \) as the default, after testing linearized multiplier accuracies obtained by running the model with and without the instability corrections. In practice \( M = 3 \) provides more than sufficient accuracy for this choice of \( \epsilon \).

Rights and permissions

Reprints and permissions

Copyright information

© 2019 Springer Nature Switzerland AG

About this chapter

Check for updates. Verify currency and authenticity via CrossMark

Cite this chapter

Spurr, R., Christi, M. (2019). The LIDORT and VLIDORT Linearized Scalar and Vector Discrete Ordinate Radiative Transfer Models: Updates in the Last 10 Years. In: Kokhanovsky, A. (eds) Springer Series in Light Scattering. Springer Series in Light Scattering. Springer, Cham. https://doi.org/10.1007/978-3-030-03445-0_1

Download citation

Publish with us

Policies and ethics