Skip to main content
Log in

Cap integration in spectral gravity forward modelling: near- and far-zone gravity effects via Molodensky’s truncation coefficients

  • Original Article
  • Published:
Journal of Geodesy Aims and scope Submit manuscript

Abstract

Spectral gravity forward modelling is a technique that converts a band-limited topography into its implied gravitational field. This conversion implicitly relies on global integration of topographic masses. In this paper, a modification of the spectral technique is presented that provides gravity effects induced only by the masses located inside or outside a spherical cap centred at the evaluation point. This is achieved by altitude-dependent Molodensky’s truncation coefficients, for which we provide infinite series expansions and recurrence relations with a fixed number of terms. Both representations are generalized for an arbitrary integer power of the topography and arbitrary radial derivative. Because of the altitude-dependency of the truncation coefficients, a straightforward synthesis of the near- and far-zone gravity effects at dense grids on irregular surfaces (e.g. the Earth’s topography) is computationally extremely demanding. However, we show that this task can be efficiently performed using an analytical continuation based on the gradient approach, provided that formulae for radial derivatives of the truncation coefficients are available. To demonstrate the new cap-modified spectral technique, we forward model the Earth’s degree-360 topography, obtaining near- and far-zone effects on gravity disturbances expanded up to degree 3600. The computation is carried out on the Earth’s surface and the results are validated against an independent spatial-domain Newtonian integration (\(1\,\upmu \mathrm {Gal}\) RMS agreement). The new technique is expected to assist in mitigating the spectral filter problem of residual terrain modelling and in the efficient construction of full-scale global gravity maps of highest spatial resolution.

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

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Fig. 1
Fig. 2
Fig. 3
Fig. 4
Fig. 5
Fig. 6
Fig. 7

Similar content being viewed by others

References

  • Balmino G (1994) Gravitational potential harmonics from the shape of an homogeneous body. Celest Mech Dyn Astron 60:331–364

    Article  Google Scholar 

  • Balmino G, Vales N, Bonvalot S, Briais A (2012) Spherical harmonic modelling to ultra-high degree of Bouguer and isostatic anomalies. J Geod 86:499–520. https://doi.org/10.1007/s00190-011-0533-4

    Article  Google Scholar 

  • Bucha B, Janák J (2014) A MATLAB-based graphical user interface program for computing functionals of the geopotential up to ultra-high degrees and orders: efficient computation at irregular surfaces. Comput Geosci 66:219–227. https://doi.org/10.1016/j.cageo.2014.02.005

    Article  Google Scholar 

  • Bucha B, Janák J, Papčo J, Bezděk A (2016) High-resolution regional gravity field modelling in a mountainous area from terrestrial gravity data. Geophys J Int 207:949–966. https://doi.org/10.1093/gji/ggw311

    Article  Google Scholar 

  • Bucha B, Hirt C, Kuhn M (2018) Runge–Krarup-type gravity field solutions to avoid divergence in traditional external spherical harmonic modelling. J Geod (submitted)

  • Comtet L (1974) Advanced combinatorics: the art of finite and infinite expansions, revised and enlarged edn. D. Reidel Publishing Company, Dordrecht, p 343

  • Eshagh M (2009) On satellite gravity gradiometry. Ph.D. thesis, Royal Institute of Technology, Division of Geodesy, Stockholm, Sweden, p 222

  • Fantino E, Casotto S (2009) Methods of harmonic synthesis for global geopotential models and their first-, second- and third-order gradients. J Geod 83:595–619. https://doi.org/10.1007/s00190-008-0275-0

    Article  Google Scholar 

  • Featherstone WE (2013) Deterministic, stochastic, hybrid and band-limited modifications of Hotine’s integral. J Geod 87:487–500. https://doi.org/10.1007/s00190-013-0612-9

    Article  Google Scholar 

  • Forsberg R (1984) A study of terrain reductions, density anomalies and geophysical inversion methods in gravity field modelling. Report No. 355, Department of Geodetic Science and Surveying, The Ohio State University, Columbus, Ohio, p 129

  • Freeden W, Schneider F (1998) Wavelet approximations on closed surfaces and their application to boundary-value problems of potential theory. Math Methods Appl Sci 21:129–163

    Article  Google Scholar 

  • Grombein T, Seitz K, Heck B (2013) Optimized formulas for the gravitational field of a tesseroid. J Geod 87:645–660. https://doi.org/10.1007/s00190-013-0636-1

    Article  Google Scholar 

  • Grombein T, Seitz K, Heck B (2017) On high-frequency topography-implied gravity signals for a height system unification using GOCE-based global geopotential models. Surv Geophys 38:443–477. https://doi.org/10.1007/s10712-016-9400-4

    Article  Google Scholar 

  • Heiskanen WA, Moritz H (1967) Physical geodesy. W. H. Freeman and Company, San Francisco, p 364

    Google Scholar 

  • Hirt C (2012) Efficient and accurate high-degree spherical harmonic synthesis of gravity field functionals at the Earth’s surface using the gradient approach. J Geod 86:729–744. https://doi.org/10.1007/s00190-012-0550-y

    Article  Google Scholar 

  • Hirt C, Kuhn M (2014) Band-limited topographic mass distribution generates full-spectrum gravity field: gravity forward modeling in the spectral and spatial domains revisited. J Geophys Res Solid Earth 119:3646–3661. https://doi.org/10.1002/2013JB010900

    Article  Google Scholar 

  • Hirt C, Kuhn M (2017) Convergence and divergence in spherical harmonic series of the gravitational field generated by high-resolution planetary topography—a case study for the Moon. J Geophys Res Planets 122:1727–1746. https://doi.org/10.1002/2017JE005298

    Article  Google Scholar 

  • Hirt C, Rexer M (2015) Earth 2014: 1 arc-min shape, topography, bedrock and ice-sheet models—available as gridded data and degree-10,800 spherical harmonics. Int J Appl Earth Obs Geoinf 39:103–112. https://doi.org/10.1016/j.jag.2015.03.001

    Article  Google Scholar 

  • Hirt C, Featherstone WE, Marti U (2010) Combining EGM2008 and SRTM/DTM2006.0 residual terrain model data to improve quasigeoid computations in mountainous areas devoid of gravity data. J Geod 84:557–567. https://doi.org/10.1007/s00190-010-0395-1

    Article  Google Scholar 

  • Hirt C, Claessens S, Fecher T, Kuhn M, Pail R, Rexer M (2013) New ultrahigh-resolution picture of Earth’s gravity field. Geophys Res Lett 40:4279–4283. https://doi.org/10.1002/grl.50838

    Article  Google Scholar 

  • Hirt C, Kuhn M, Claessens S, Pail R, Seitz K, Gruber T (2014) Study of the Earth’s short-scale gravity field using the ERTM2160 gravity model. Comput Geosci 73:71–80. https://doi.org/10.1016/j.cageo.2014.09.001

    Article  Google Scholar 

  • Hirt C, Reußner E, Rexer M, Kuhn M (2016) Topographic gravity modeling for global Bouguer maps to degree 2160: validation of spectral and spatial domain forward modeling techniques at the 10 microgal level. J Geophys Res Solid Earth 121:6846–6862. https://doi.org/10.1002/2016JB013249

    Article  Google Scholar 

  • Hoffmann-Wellenhof B, Moritz H (2005) Physical Geodesy. Springer, New York, p 403

    Google Scholar 

  • Holmes SA (2003) High degree spherical harmonic synthesis for simulated earth gravity modelling. Ph.D. thesis, Department of Spatial Sciences, Curtin University of Technology, Perth, Australia, p 171

  • Kuhn M, Hirt C (2016) Topographic gravitational potential up to second-order derivatives: an examination of approximation errors caused by rock-equivalent topography (RET). J Geod 90:883–902. https://doi.org/10.1007/s00190-016-0917-6

    Article  Google Scholar 

  • Makhloof AA, Ilk KH (2008) Far-zone effects for different topographic-compensation models based on a spherical harmonic expansion of the topography. J Geod 82:613–635. https://doi.org/10.1007/s00190-008-0214-0

    Article  Google Scholar 

  • Martinec Z (1998) Boundary-value problems for gravimetric determination of a precise geoid. Springer, Berlin, p 223

    Google Scholar 

  • Mikuška J, Pašteka R, Marušiak I (2006) Estimation of distant relief effect in gravimetry. Geophysics 71:J59–J69. https://doi.org/10.1190/1.2338333

    Article  Google Scholar 

  • Moazezi S, Zomorrodian H, Siahkoohi HR, Azmoudeh-Ardalan A, Gholami A (2016) Fast ultrahigh-degree global spherical harmonic synthesis on nonequispaced grid points at irregular surfaces. J Geod 90:853–870. https://doi.org/10.1007/s00190-016-0915-8

    Article  Google Scholar 

  • Molodensky MS, Eremeev VF, Yurkina MI (1962) Methods for study of the external gravitational field and figure of the Earth. Israel Program for Scientific Translations, Jerusalem, p 248, translated from Russian (1960)

  • Moreaux G, Tscherning CC, Sanso F (1999) Approximation of harmonic covariance functions on the sphere by non-harmonic locally supported functions. J Geod 73:555–567

    Article  Google Scholar 

  • Nagy D, Papp G, Benedek J (2000) The gravitational potential and its derivatives for the prism. J Geod 74:552–560

    Article  Google Scholar 

  • Novák P, Vaníček P, Martinec Z, Véronneau M (2001) Effects of the spherical terrain on gravity and the geoid. J Geod 75:491–504

    Article  Google Scholar 

  • Paul MK (1973) A method of evaluating the truncation error coefficients for geoidal height. Bull Géod 110:413–425

    Article  Google Scholar 

  • Paul MK (1983) Recurrence relations for the truncation error coefficients for the extended Stokes function. Bull Géod 57:152–166

    Article  Google Scholar 

  • Pavlis NK (1991) Estimation of geopotential differences over intercontinental locations using satellite and terrestrial measurements. Report No. 409, Department of Geodetic Science, The Ohio State University, Ohio, USA, p 155

  • Pohánka V (1988) Optimum expression for computation of the gravity field of a homogeneous polyhedral body. Geophys Prospect 35:733–751

    Article  Google Scholar 

  • Rexer M, Hirt C (2015) Ultra-high-degree surface spherical harmonic analysis using the Gauss–Legendre and the Driscoll/Healy quadrature theorem and application to planetary topography models of Earth, Mars and Moon. Surv Geophys 36:803–830. https://doi.org/10.1007/s10712-015-9345-z

    Article  Google Scholar 

  • Rexer M, Hirt C, Bucha B, Holmes S (2018) Solution to the spectral filter problem of residual terrain modelling RTM. J Geod. https://doi.org/10.1007/s00190-017-1086-y

    Google Scholar 

  • Shepperd SW (1982) A recursive algorithm for evaluating Molodenskii-type truncation error coefficients at altitude. Bull Géod 56:95–105

    Article  Google Scholar 

  • Sjöberg LE (2005) A discussion on the approximations made in the practical implementation of the remove-compute-restore technique in regional geoid modelling. J Geod 78:645–653. https://doi.org/10.1007/s00190-004-0430-1

    Article  Google Scholar 

  • Sneeuw N (1994) Global spherical harmonic analysis by least-squares and numerical quadrature methods in historical perspective. Geophys J Int 118:707–716

    Article  Google Scholar 

  • Šprlák M, Hamáčková E, Novák P (2015) Alternative validation method of satellite gradiometric data by integral transform of satellite altimetry data. J Geod 89:757–773. https://doi.org/10.1007/s00190-015-0813-5

    Article  Google Scholar 

  • Tenzer R, Novák P, Vajda P, Ellmann A, Abdalla A (2011) Far-zone gravity field contributions corrected for the effect of topography by means of Molodensky’s truncation coefficients. Stud Geophys Geod 55:55–71

    Article  Google Scholar 

  • Thalhammer M (1995) Regionale Gravitationsfeldbestimmung mit zukünftigen Satellitenmissionen (SST und Gradiometrie). Reihe C, Heft Nr. 437, Deutsche Geodätische Kommission, München, Germany, p 96 (in German)

  • Wessel P, Smith WHF (1998) New, improved version of generic mapping tools released. EOS Trans Am Geophys Union 79:579. https://doi.org/10.1029/98EO00426

    Article  Google Scholar 

  • Wieczorek MA, Phillips RJ (1998) Potential anomalies on a sphere: applications to the thickness of the lunar crust. J Geophys Res 103:1715–1724

    Article  Google Scholar 

Download references

Acknowledgements

Blažej Bucha was supported by the ProjectVEGA 1/0954/15 and acknowledges the computational resources made available by the HPC centres at the Slovak University of Technology in Bratislava and at the Slovak Academy of Sciences, which are parts of the Slovak Infrastructure of High Performance Computing (SIVVP project, ITMS code 26230120002, funded by the European region development funds, ERDF). Christian Hirt would like to thank the German National Research Foundation (DFG) for providing funding under Grant Agreement Hi 1760/01. The spatial-domain Newtonian integration was performed using the supercomputing resources kindly provided by Western Australia’s Pawsey Supercomputing Center. The maps were produced using the Generic Mapping Tools (Wessel and Smith 1998).

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Blažej Bucha.

Appendices

Appendix A: Derivation of cap-modified spectral gravity forward modelling

In this appendix, we derive Eq. (6) that yields the gravitational effect induced by topographic masses located inside/outside a spherical cap defined by the spherical distance \(\psi _0\) from the computation point. To this end, we closely follow the derivations described, for instance, in Section 7–4 of Heiskanen and Moritz (1967). At first, let us focus on the near-zone effect.

Starting from the spherical harmonic expansion as provided, for instance, by Wieczorek and Phillips (1998), the topographic potential induced by topographic masses all around the globe can be written as

$$\begin{aligned} V(r,\varphi ,\lambda )= & {} 2\pi \, G \rho R^2 \sum _{n=0}^{\infty } \left( \frac{R}{r} \right) ^{n+1}\nonumber \\&\times \,\frac{2}{2n+1} \, \sum _{p=1}^{n+3}\frac{\prod \nolimits _{i=1}^{p}(n+4-i)}{p! \, (n+3)} \, H_n^p(\varphi ,\lambda ) \, ,\nonumber \\ \end{aligned}$$
(19)

with \(H_n^p(\varphi ,\lambda )\) being the Laplace surface spherical harmonic of the topographic height function (see Eq. 5),

$$\begin{aligned} H_n^p(\varphi ,\lambda ) = \sum _{m=-n}^{n} \bar{H}_{nmp} \, \bar{Y}_{nm}(\varphi ,\lambda ) \, . \end{aligned}$$
(20)

The term \(H_n^p(\varphi ,\lambda )\) can equally be written as

$$\begin{aligned} H_n^p(\varphi ,\lambda )= & {} \frac{2n+1}{4\pi }\nonumber \\&\times \int \limits _{\psi =0}^{\pi } \int \limits _{\alpha =0}^{2\pi } H^p(\psi ,\alpha ) \, P_n(\cos \psi ) \, \sin \psi \, \mathrm d\alpha \, \mathrm d\psi \, ,\nonumber \\ \end{aligned}$$
(21)

where \(P_n\) is unnormalized Legendre polynomial of degree n, and \(\psi \) and \(\alpha \) are, respectively, the spherical distance and the azimuth between the computation point and the integration element. Substituting Eq. (21) into Eq. (19) and interchanging the order of summations and integrations, we get

$$\begin{aligned} V(r,\varphi ,\lambda )= & {} G \rho R^2 \sum _{p=1}^{\infty } \int \limits _{\psi =0}^{\pi } \int \limits _{\alpha =0}^{2\pi } H^p(\psi ,\alpha ) K_p(r,\psi )\nonumber \\&\times \sin \psi \, \mathrm d\alpha \, \mathrm d\psi \, , \end{aligned}$$
(22)

where

$$\begin{aligned} K_p(r,\psi ) = \sum _{n=0}^{\infty } \left( \frac{R}{r} \right) ^{n+1} \, \frac{\prod \nolimits _{i=1}^{p}(n+4-i)}{p! \, (n+3)} \, P_n(\cos \psi ) \, .\nonumber \\ \end{aligned}$$
(23)

Such reordering of integration and summation is permissible as long as the series is uniformly convergent. This holds true for points with \(r > \max (R+\hat{H}(\varphi ,\lambda ))\) (e.g. Balmino 1994; Wieczorek and Phillips 1998). Note that to be the series in Eq. (23) convergent, the condition \(r > R\) must be satisfied (e.g. Martinec 1998).

Restricting the integration limits over \(\psi \) in Eq. (22) from \(\psi \in [0,\, \pi ]\) (global integration) to \(\psi \in [0,\, \psi _0]\) (inside-cap integration) yields the near-zone contribution of topographic masses to the topographic potential,

$$\begin{aligned} V^\mathrm {In}(r,\varphi ,\lambda )= & {} G \rho R^2 \sum _{p=1}^{\infty } \int \limits _{\psi =0}^{\psi _0} \int \limits _{\alpha =0}^{2\pi } H^p(\psi ,\alpha ) \, K_p(r,\psi )\nonumber \\&\times \sin \psi \, \mathrm d\alpha \, \mathrm d\psi \, . \end{aligned}$$
(24)

After introducing the discontinuous function

$$\begin{aligned} K^{\mathrm {In}}_p(r,\psi )= {\left\{ \begin{array}{ll} K_p(r,\psi ) \quad &{}\mathrm {for} \quad 0\le \psi \le \psi _0 \, ,\\ 0 \quad &{}\mathrm {for} \quad \psi _0<\psi \le \pi \, , \end{array}\right. } \end{aligned}$$
(25)

Eq. (24) can be expressed as a global integration of the form

$$\begin{aligned} V^\mathrm {In}(r,\varphi ,\lambda )= & {} G \rho R^2 \sum _{p=1}^{\infty } \int \limits _{\psi =0}^{\pi } \int \limits _{\alpha =0}^{2\pi } H^p(\psi ,\alpha ) \, K^{\mathrm {In}}_p(r,\psi )\nonumber \\&\times \sin \psi \, \mathrm d\alpha \, \mathrm d\psi \, . \end{aligned}$$
(26)

Next, the kernel function \(K^{\mathrm {In}}_p(r,\psi )\) can be expanded in an infinite series of Legendre polynomials

$$\begin{aligned} K^{\mathrm {In}}_p(r,\psi ) = \sum _{n=0}^{\infty } \frac{2n+1}{2} \, Q_{np}^{\mathrm {In}}(r,\psi _0) \, P_n(\cos \psi ) \end{aligned}$$
(27)

with the truncation coefficients \(Q_{np}^{\mathrm {In}}(r,\psi _0)\) given by

$$\begin{aligned} Q_{np}^{\mathrm {In}}(r,\psi _0) = \int \limits _{0}^{\pi } K^{\text {In}}_p(r,\psi ) \, P_n(\cos \psi ) \, \sin \psi \, \mathrm d\psi \, . \end{aligned}$$
(28)

In Appendices B and C, respectively, we provide spectral and recurrence relations suitable to evaluate numerically the truncation coefficients \(Q_{np}^{\mathrm {In}}(r,\psi _0)\) as well as \(Q_{np}^{\mathrm {Out}}(r,\psi _0)\).

Substituting Eq. (27) into (26) and with the help of Eq. (21), we obtain

$$\begin{aligned} V^\mathrm {In}(r,\varphi ,\lambda ) = 2\pi \,G \rho R^2 \sum _{p=1}^{\infty } \sum _{n=0}^{\infty } Q_{np}^{\mathrm {In}}(r,\psi _0) \, H_n^p(\varphi ,\lambda ) \, .\nonumber \\ \end{aligned}$$
(29)

Finally, realizing that (cf. Lemma 4.1 of Freeden and Schneider 1998)

$$\begin{aligned} H_n^p(\varphi ,\lambda ) = 0 \quad \mathrm {for} \quad n>p\times n_{\max }, \end{aligned}$$
(30)

then utilizing Eq. (20) and, finally, after truncating the infinite series over p at some \(p_{\max }\), we arrive at Eq. (6) for \(z=\text {`In'}\).

The far-zone gravitational effect \(V^\mathrm {Out}(r,\varphi ,\lambda )\) from Eq. (6), z being equal to \(\text {`Out'}\), can be derived in a similar manner by integrating over \(\psi \in [\psi _0,\, \pi ]\). These derivations require to introduce the discontinuous function

$$\begin{aligned} K^{\mathrm {Out}}_p(r,\psi )= {\left\{ \begin{array}{ll} 0 \quad &{}\mathrm {for} \quad 0\le \psi < \psi _0 \, ,\\ K_p(r,\psi ) \quad &{}\mathrm {for} \quad \psi _0\le \psi \le \pi \, , \end{array}\right. } \end{aligned}$$
(31)

which is then utilized in the definition of the far-zone truncation coefficients

$$\begin{aligned} Q_{np}^{\mathrm {Out}}(r,\psi _0) = \int \limits _{0}^{\pi } K^{\text {Out}}_p(r,\psi ) \, P_n(\cos \psi ) \, \sin \psi \, \mathrm d\psi \, . \end{aligned}$$
(32)

Appendix B: Spectral representation of truncation coefficients

The spectral form of the near-zone truncation coefficients can be obtained by substituting Eqs. (23) and (25) into Eq. (28). Then, after interchanging the order of summation and integration, the near-zone truncation coefficients are given by

$$\begin{aligned} Q_{np}^{\mathrm {In}}(r,\psi _0)= & {} \sum _{l=0}^{\infty } \left( \frac{R}{r} \right) ^{l+1} \, \frac{\prod \nolimits _{i=1}^{p}(l+4-i)}{p! \, (l+3)}\nonumber \\&\times \int \limits _{0}^{\psi _0} P_l(\cos \psi ) \, P_n(\cos \psi ) \, \sin \psi \, \mathrm d\psi \, . \end{aligned}$$
(33)

The integrals

$$\begin{aligned} I_{ln}(u_0,1)= & {} \int \limits _{0}^{\psi _0} P_l(\cos \psi ) \, P_n(\cos \psi ) \, \sin \psi \, \mathrm d\psi \nonumber \\= & {} \int \limits _{u_0}^{1} P_l(u) \, P_n(u) \, \mathrm du \, , \end{aligned}$$
(34)

with the substitutions \(u=\cos \psi \) and \(u_0=\cos \psi _0\), can be evaluated numerically via recurrence relations. For instance, Paul (1973) provides recurrence relations to compute

$$\begin{aligned} I_{ln}(-1,u_0) = \int \limits _{-1}^{u_0} P_l(u) \, P_n(u) \, \mathrm du \, , \end{aligned}$$
(35)

from which we can easily obtain \(I_{ln}(u_0,1)\) by utilizing the orthogonality property of Legendre polynomials

$$\begin{aligned} I_{ln}(-1,1) = \frac{2}{2n+1} \delta _{ln} \, , \end{aligned}$$
(36)

where

$$\begin{aligned} \delta _{ln} = {\left\{ \begin{array}{ll} 1\,, \quad \quad l=n \, ,\\ 0\,, \quad \quad l\ne n \, . \end{array}\right. } \end{aligned}$$
(37)

The integrals \(I_{ln}(u_0,1)\) thus read

$$\begin{aligned} I_{ln}(u_0,1)= & {} I_{ln}(-1,1) - I_{ln}(-1,u_0)\nonumber \\= & {} \frac{2}{2n+1} \delta _{ln} - I_{{ln}}(-1,u_0) \, . \end{aligned}$$
(38)

While Eq. (35) is needed for far-zone effects, Eq. (38) enters the computation of the near-zone gravity contribution. More general recurrence relations to compute \(I_{ln}(u_1,u_2)\), suitable both for near- and far-zone effects, were derived by Moreaux et al. (1999).

Needless to say, the far-zone truncation coefficients can be derived by substituting Eqs. (23) and (31) into Eq. (32), obtaining

$$\begin{aligned} Q_{np}^{\mathrm {Out}}(r,\psi _0)= & {} \sum _{l=0}^{\infty } \left( \frac{R}{r} \right) ^{l+1} \, \frac{\prod \nolimits _{i=1}^{p}(l+4-i)}{p! \, (l+3)}\nonumber \\&\times \int \limits _{\psi _0}^{\pi } P_l(\cos \psi ) \, P_n(\cos \psi ) \, \sin \psi \, \mathrm d\psi \, . \end{aligned}$$
(39)

Appendix C: Recurrence representation of truncation coefficients

In this appendix, we derive recurrence relations with a fixed number of terms to compute the truncation coefficients for an arbitrary n and p.

From Martinec (1998) it can be seen that the integral kernels \(K_p(r,\psi )\) from Eq. (23) have the following closed spatial forms

$$\begin{aligned} K_1(r,\psi )= & {} \, \frac{R}{l(r,\psi )} \, , \nonumber \\ K_2(r,\psi )= & {} \frac{1}{2} \left( K_1(r,\psi ) - r \, \frac{\partial K_1(r,\psi )}{\partial r} \right) \, ,\nonumber \\ K_p(r,\psi )= & {} \frac{1}{p!} \sum _{s=1}^{p-2} a_{ps}\, r^{p-s} \, \frac{\partial ^{p-s} K_1(r,\psi )}{\partial r^{p-s}}\, , \quad p \ge 3\, ,\nonumber \\ \end{aligned}$$
(40)

with the Euclidean distance

$$\begin{aligned} l(r,\psi )=\sqrt{r^2-2 R r \cos \psi +R^2} \end{aligned}$$
(41)

and the coefficients

$$\begin{aligned} a_{ps}=(-1)^{p-1} \frac{(p-1)! \, (p-3)!}{(p-s)! \, (p-s-2)! \, (s-1)!} \, . \end{aligned}$$
(42)

After introducing the substitution \(t=R \slash r\) and the normalized Euclidean distance

$$\begin{aligned} g(t,u)=\sqrt{1-2\, tu+t^2}\, , \end{aligned}$$
(43)

we obtain from Eqs. (28) and (32) with the help of Eq. (40)

$$\begin{aligned} Q_{n1}(r,\psi _0)= & {} G_n^{(0)}(t,u_0) \, , \nonumber \\ Q_{n2}(r,\psi _0)= & {} \frac{1}{2} \left( G_n^{(0)}(t,u_0) - r \, G_n^{(1)}(t,u_0) \right) \, ,\nonumber \\ Q_{np}(r,\psi _0)= & {} \frac{1}{p!} \sum _{s=1}^{p-2} a_{ps}\, r^{p-s} G_n^{(p-s)}(t,u_0)\, , \quad p \ge 3\, .\nonumber \\ \end{aligned}$$
(44)

In Eq. (44), we introduced the substitution

$$\begin{aligned} G_n^{(j)}(t,u_0) = \frac{\partial ^j}{\partial r^j} L_n(t,u_0) \, , \quad j \ge 0\, , \end{aligned}$$
(45)

with

$$\begin{aligned} L_n(t,u_0) = t\, A_n(t,u_0) \end{aligned}$$
(46)

and

$$\begin{aligned} A_n(t,u_0) = \int \limits _{u_l}^{u_u} \frac{P_n(u)}{g(t,u)} \, \mathrm du \, , \end{aligned}$$
(47)

where \(u_u\) and \(u_l\) denote the upper and lower integration limits, respectively. For the computation of \(Q^{\mathrm {In}}_{np}(r,\psi _0)\), we thus have \(u_u=1\) and \(u_l=u_0\). For \(Q^{\mathrm {Out}}_{np}(r,\psi _0)\), the limits are \(u_u=u_0\) and \(u_l=-\,1\). For simplicity, the superscript z is omitted in Eq. (44), because these relations hold both for \(Q_{np}^{\mathrm {In}}(r,\psi _0)\) and \(Q_{np}^{\mathrm {Out}}(r,\psi _0)\), depending on \(u_u\) and \(u_l\).

To compute \(G^{(j)}_n(t,u_0)\), representing the jth derivative of the composite function \(L_n(t,u_0)\) in terms of r, we use the chain rule generalized for an arbitrarily high derivative (Faà di Bruno’s formula; e.g. Comtet 1974). This rule, expressed via the partial Bell polynomials \(\tilde{B}_{jk}\), reads for \(j\ge 0\)

$$\begin{aligned} G^{(j)}_n(t,u_0)= & {} \sum _{k=0}^{j} J_n^{(k)}(t,u_0)\nonumber \\&\times \tilde{B}_{jk} \left( T^{(1)}(r),T^{(2)}(r), \dots , T^{(j-k+1)}(r) \right) \nonumber \\ \end{aligned}$$
(48)

with

$$\begin{aligned} J_n^{(v)}(t,u_0) = \frac{\partial ^v}{\partial t^v} L_n(t,u_0)\,, \quad v \ge 0\, , \end{aligned}$$
(49)

and

$$\begin{aligned} T^{(v)}(r) = \frac{\mathrm d^v t}{\mathrm dr^v} = (-1)^{v} \, v! \, \frac{R}{r^{v+1}}\, ,\quad v \ge 0\, . \end{aligned}$$
(50)

The partial Bell polynomials \(\tilde{B}_{jk}\) can be computed recursively by the relation (e.g. Comtet 1974)

$$\begin{aligned} \tilde{B}_{jk} = \sum _{s=1}^{j-k+1} \begin{pmatrix} j-1\\ s-1 \end{pmatrix} T^{(s)} \, \tilde{B}_{j-s,k-1} \, , \quad j,k\ge 1\, , \end{aligned}$$
(51)

with

$$\begin{aligned} \tilde{B}_{00}= & {} 1 \, ,\nonumber \\ \tilde{B}_{j0}= & {} 0\,, \quad j\ge 1 \, ,\nonumber \\ \tilde{B}_{0k}= & {} 0\,, \quad k\ge 1 \, . \end{aligned}$$
(52)

Equation (49) implies that an arbitrary derivative of a product of t and \(A_n(t,u_0)\) (cf. Eq. 46) has to be found. This can be done using the general Leibniz rule. Having two n-times differentiable functions \(f_1(x)\) and \(f_2(x)\), the rule says that the nth derivative of the product \(f_1(x)\, f_2(x)\) is given as (e.g. Comtet 1974)

$$\begin{aligned} \frac{\mathrm d^n f_1(x)\, f_2(x)}{\mathrm dx^n} = \sum _{k=0}^{n} \begin{pmatrix} n\\ k \end{pmatrix} \frac{\mathrm d^{n-k} f_1(x)}{\mathrm dx^{n-k}} \, \frac{\mathrm d^{k} f_2(x)}{\mathrm dx^{k}} \, . \end{aligned}$$
(53)

Setting \(f_1=t\) and \(f_2=A_n(t,u_0)\) and after some simplifications, we find that the derivatives of \(L_n(t,u_0)\) with respect to t read

$$\begin{aligned} J_n^{(v)}(t,u_0)= & {} \frac{\partial ^v}{\partial t^v} L_n(t,u_0)\nonumber \\= & {} t\, \frac{\partial ^v}{\partial t^v} A_n(t,u_0) + v \, \frac{\partial ^{v-1}}{\partial t^{v-1}} A_n(t,u_0) \, , \quad v\ge 1\, .\nonumber \\ \end{aligned}$$
(54)

This, however, requires to differentiate \(A_n(t,u_0)\) with respect to t,

$$\begin{aligned} A_{n}^{(i)}(t,u_0) = \frac{\partial ^i}{\partial t^i} A_n(t,u_0) = \frac{\partial ^i}{\partial t^i} \int \limits _{u_l}^{u_u} \frac{P_n(u)}{g(t,u)} \, \mathrm du\, , \quad i\ge 0\, .\nonumber \\ \end{aligned}$$
(55)

Therefore, from the recurrence relations for \(A^{(0)}_n(t,u_0)\) provided by Pavlis (1991), we have derived relations for \(A_{n}^{(i)}(t,u_0)\), \(i\ge 0\):

$$\begin{aligned} A_0^{(i)}(t,u_0)= & {} - \left[ \frac{\partial ^i}{\partial t^i} \alpha _0(t,u)\, \gamma (t,u) \right] _{u_l}^{u_u}\,, \end{aligned}$$
(56)
$$\begin{aligned} A_1^{(i)}(t,u_0)= & {} -\frac{1}{3} \left[ \frac{\partial ^i}{\partial t^i} \alpha _1(t,u)\, \gamma (t,u) \right] _{u_l}^{u_u}\,, \end{aligned}$$
(57)
$$\begin{aligned} A_n^{(i)}(t,u_0)= & {} \frac{1}{2n+1} \Bigg [ -(t^2+1) B_n^{(i)}(t,u_0)\nonumber \\&+\, 2t\,B_{n-1}^{(i)}(t,u_0) - 2i\, t \, B_{n}^{(i-1)}(t,u_0)\nonumber \\&+\,2i\,B_{n-1}^{(i-1)}(t,u_0) - i(i-1) B_{n}^{(i-2)}(t,u_0)\nonumber \\&+\,2(n+1)\, M_n(u_0) \, \gamma ^{(i)}(t,u_0) \Bigg ]\, , \quad n\ge 2\, ,\nonumber \\ \end{aligned}$$
(58)

where the ith derivatives of \(\alpha _0\), \(\alpha _1\) and \(\gamma \) with respect to t read

$$\begin{aligned} \alpha _0^{(i)}(t,u)= & {} {\left\{ \begin{array}{ll} \begin{aligned} &{}t+\frac{1}{t}-2u\,, \quad &{}&{}i=0\,,\\ &{}(-1)^i \, \frac{i!}{t^{i+1}} + \delta _{\alpha _0}^{(i)}\,, \quad &{}&{}i\ge 1\,, \end{aligned} \end{array}\right. } \end{aligned}$$
(59)
$$\begin{aligned} \alpha _1^{(i)}(t,u)= & {} {\left\{ \begin{array}{ll} \begin{aligned} &{}t^2-tu-\frac{u}{t}+\frac{1}{t^2}-2u^2+2\,, \quad &{}&{} i=0\,,\\ &{}(-1)^{i+1} \, \frac{i!\, u}{t^{i+1}}\\ &{}\quad + (-1)^{i} \, \frac{(i+1)!}{t^{i+2}} + \delta _{\alpha _1}^{(i)}\, , \quad &{}&{}i\ge 1\,, \end{aligned} \end{array}\right. } \end{aligned}$$
(60)
$$\begin{aligned} \gamma ^{(i)}(t,u)= & {} {\left\{ \begin{array}{ll} \begin{aligned} &{}\frac{1}{g(t,u)}\,, &{}&{} i=0\,,\\ &{} \frac{\partial ^i}{\partial t^i} \frac{1}{g(t,u)} = \sum _{ \begin{matrix} (s+i)\ \text {is even}\\ s=0 \end{matrix} }^{i} (-1)^{\frac{i+s}{2}}\\ &{} \times \frac{(i-s+1)!!\, (i+s-1)!!}{(i-s+1)!}&{}&{}\\ &{}\times \frac{i!}{s!} \, \frac{(t-u)^s}{g^{i+s+1}(t,u)} \, ,\quad &{}&{}i\ge 1\,, \end{aligned} \end{array}\right. } \end{aligned}$$
(61)

with

$$\begin{aligned} \delta _{\alpha _0}^{(i)}= {\left\{ \begin{array}{ll} \begin{aligned} &{}1\,, \quad &{}&{} i=1\,, \\ &{}0\,, \quad &{}&{} i\ge 2\,, \end{aligned} \end{array}\right. } \end{aligned}$$
(62)

and

$$\begin{aligned} \delta _{\alpha _1}^{(i)}= {\left\{ \begin{array}{ll} \begin{aligned} &{}2t-u\,, \quad &{}&{} i=1\,, \\ &{}2\,, \quad &{}&{} i=2\,,\\ &{}0\,, \quad &{}&{} i\ge 3\,.\\ \end{aligned} \end{array}\right. } \end{aligned}$$
(63)

After simple manipulations, it can be shown that Eqs. (56) and (57) for \(i=0\) are equal to Eq. (A.30b) of Pavlis (1991). By differentiating with respect to t, we obtained Eqs. (56) and (57) for \(i\ge 1\). For their numerical evaluation, we use the general Leibniz rule (Eq. 53). Similarly, Eq. (58) for \(i\ge 1\) can be derived by differentiating the relation for \(i=0\), which is provided by Pavlis (1991) in Eq. (A.30a). Note that in Eq. (58), the negative derivatives that occur for \(i=0,1\) need to be set to zero. Finally, Eq. (61) for \(i\ge 1\) follows from Eq. (A.5.12) of Martinec (1998).

Analogously, the recurrence relations for \(B_n^{(i)}(t,u_0)\), \(i\ge 0\), (see Eq. 58) were derived from the formulae for \(B_n^{(0)}(t,u_0)\) that are provided by Pavlis (1991):

$$\begin{aligned} B_0^{(i)}(t,u_0)= & {} \left[ \frac{\partial ^i}{\partial t^i} \beta _0(t,u)\, \gamma (t,u) \right] _{u_l}^{u_u}\,, \end{aligned}$$
(64)
$$\begin{aligned} B_1^{(i)}(t,u_0)= & {} \left[ \frac{\partial ^i}{\partial t^i} \beta _1(t,u)\, \gamma (t,u) \right] _{u_l}^{u_u}\,, \end{aligned}$$
(65)
$$\begin{aligned} B_n^{(i)}(t,u_0)= & {} \frac{t^2+1}{t} B_{n-1}^{(i)}(t,u_0) - B_{n-2}^{(i)}(t,u_0)\nonumber \\&-\,\frac{i}{t} B_{n}^{(i-1)}(t,u_0) + 2i B_{n-1}^{(i-1)}(t,u_0)\nonumber \\&-\,\frac{i}{t} B_{n-2}^{(i-1)}(t,u_0)+\frac{i(i-1)}{t} B_{n-1}^{(i-2)}(t,u_0)\nonumber \\&-\,\frac{1}{t}\, M_{n-1}(u_0)\, \gamma ^{(i)}(t,u_0)\, , \quad n\ge 2\, , \end{aligned}$$
(66)

where the ith derivatives of \(\beta _0\) and \(\beta _1\) with respect to t are given as

$$\begin{aligned} \beta _0^{(i)}= & {} (-1)^i\, i! \frac{1}{t^{i+1}}\, , \quad i\ge 0\,, \end{aligned}$$
(67)
$$\begin{aligned} \beta _1^{(i)}= & {} {\left\{ \begin{array}{ll} \begin{aligned} &{}\frac{1}{t^2}-\frac{u}{t}+1\, , \quad &{}&{}i=0\,,\\ &{}(-1)^i\, (i+1)! \frac{1}{t^{i+2}} + (-1)^{i+1}\, i! \, \frac{u}{t^{i+1}}\, , \quad &{}&{}i\ge 1\,, \end{aligned} \end{array}\right. } \end{aligned}$$
(68)

Starting from Eqs. (64) and (65) for \(i=0\) (Eq. A.29b of Pavlis 1991), the relations for \(i\ge 1\) are derived by differentiation with respect to t. Again, we evaluate these equations by the general Leibniz rule (Eq. 53). The recurrence relation in Eq. (66) for \(i\ge 1\) was derived by differentiating Eq. (66) for \(i=0\) (Eq. A.29a of Pavlis 1991) with respect to t. Setting the negative derivatives to zero, Eq. (66) holds for an arbitrary \(i\ge 0\).

For the term \(M_n(t,u_0)\), Pavlis (1991) provides recurrence relations with the initial values \(u_l=-1\) and \(u_u=u_0\) (the far-zone effect in terms of the truncation coefficients). After simple manipulations, these initial values can be rewritten in a unified manner both for near- and far-zone effects, obtaining

$$\begin{aligned} M_0(u_0)= & {} [P_{1}(u)]_{u_l}^{u_u}\,, \end{aligned}$$
(69)
$$\begin{aligned} M_1(u_0)= & {} \frac{1}{3} \left[ P_2(u) - P_0(u)\right] _{u_l}^{u_u}\,, \end{aligned}$$
(70)
$$\begin{aligned} M_n(u_0)= & {} \frac{1}{n+1} [(2n-1) \, u_0 \, M_{n-1}(u_0)\nonumber \\&- (n-2) \, M_{n-2}(u_0)] \, , \quad n\ge 2\, , \end{aligned}$$
(71)

where the Legendre polynomials \(P_n(u)\) for \(n=0,1,2\) read

$$\begin{aligned} P_0(u)= & {} 1\, ,\nonumber \\ P_1(u)= & {} u\, ,\nonumber \\ P_2(u)= & {} -\,\frac{1}{2} + \frac{3}{2} \, u^2\, . \end{aligned}$$
(72)

As for the proofs of the newly derived Eqs. (58) and (66) for \(i\ge 1\), they are omitted here for the sake of brevity, but the validity of these recurrence relations can be proved using mathematical induction.

Our last remark concerns Eq. (47). Following the derivations by Pavlis (1991), recurrence relations to compute directly \(A_n(r,\psi _0)\) could possibly be obtained. As an advantage, Faà di Bruno’s formula could be avoided (cf. Eqs. 45 and 48), thus slightly simplifying the computations. However, though not confirmed by numerical experiments, an increased numerical instability can be expected when working with the Euclidean distance \(l(r,\psi )\) (Eq. 41) instead of its normalized counterpart g(tu) (Eq. 43). This is the reason why we prefer to work with \(A_n(t,u_0)\), even at the cost of somewhat increased complexity of the computation.

Appendix D: Spectral and recurrence relations for an arbitrary radial derivative of truncation coefficients

Differentiating Eq. (33) k times with respect to r, \(k\ge 1\), leads to spectral relations

$$\begin{aligned} \frac{\partial ^k}{\partial r^k}Q_{np}^{\mathrm {In}}(r,\psi _0)= & {} \frac{(-1)^k}{R^k} \sum _{l=0}^{\infty } \prod \limits _{j=1}^{k}(l+j)\nonumber \\&\times \left( \frac{R}{r} \right) ^{l+k+1} \, \frac{\prod \limits _{i=1}^{p}(l+4-i)}{p! \, (l+3)}\nonumber \\&\times \int \limits _{0}^{\psi _0} P_l(\cos \psi ) \, P_n(\cos \psi ) \, \sin \psi \, \mathrm d\psi \, .\nonumber \\ \end{aligned}$$
(73)

A similar equation for \(\partial ^k (Q_{np}^{\mathrm {Out}}(r,\psi _0)) \slash \partial r^k\), \(k\ge 1\), can analogously be obtained from Eq. (39). Let us recall that the integrals can be computed analytically (cf. Appendix B).

The recurrence relations, holding both for near- and far-zone effects, were derived by differentiating Eq. (44) with respect to r. After considering Eq. (45), we get for \(k\ge 0\)

$$\begin{aligned} \frac{\partial ^k}{\partial r^k}Q_{n1}(r,\psi _0)= & {} G_n^{(k)}(t,u_0) \, , \nonumber \\ \frac{\partial ^k}{\partial r^k}Q_{n2}(r,\psi _0)= & {} \frac{1}{2} \left( -(k-1) G_n^{(k)}(t,u_0) - r \, G_n^{(k+1)}(t,u_0) \right) \, ,\nonumber \\ \frac{\partial ^k}{\partial r^k}Q_{np}(r,\psi _0)= & {} \frac{1}{p!} \sum _{s=1}^{p-2} a_{ps} \sum _{i=0}^{k} \begin{pmatrix} k\\ i \end{pmatrix}\nonumber \\&\times R_{p-s}^{(k-i)}(r)\,G_n^{(p-s+i)}(t,u_0)\, , \quad p \ge 3\, , \end{aligned}$$
(74)

with

$$\begin{aligned} R_w^{(v)}(r) = \frac{\mathrm d^v}{\mathrm dr^v} r^w = {\left\{ \begin{array}{ll} \begin{aligned} &{}r^w\, , \ \ &{}&{}v=0 \, ,\, w\ge 1\,,\\ &{}\prod _{j=1}^{v} (w-j+1)\, r^{w-v}\, , \ \ &{}&{}v\ge 1 \,,\, w\ge 1. \end{aligned} \end{array}\right. } \end{aligned}$$
(75)

In Eq. (74), we applied the general Leibniz rule (Eq. 53) in the third case for \(p\ge 3\).

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Bucha, B., Hirt, C. & Kuhn, M. Cap integration in spectral gravity forward modelling: near- and far-zone gravity effects via Molodensky’s truncation coefficients. J Geod 93, 65–83 (2019). https://doi.org/10.1007/s00190-018-1139-x

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00190-018-1139-x

Keywords

Navigation