Skip to main content
Log in

Characterization and stabilization of the downward continuation problem for airborne gravity data

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

Abstract

In this study, we compare six commonly used methods for the downward continuation of airborne gravity data. We consider exact and noisy simulated data on grids and along flight trajectories and real data from the GRAV-D airborne campaign. We use simulated and real surface gravity data for validation. The methods comprise spherical harmonic analysis, least-squares collocation, residual least-squares collocation, least-squares radial basis functions, the inverse Poisson method and Moritz’s analytical downward continuation method. We show that all the methods perform similar in terms of surface gravity values. For real data, the downward continued airborne gravity values are used to compute a geoid model using a Stokes-integral-based approach. The quality of the computed geoid model is validated using high-quality GSVS17 GPS-levelling data. We show that the geoid model quality is similar for all the methods. However, the least-squares collocation approach appears to be more flexible and easier to use than the other methods provided that the optimal covariance function is found. We recommend it for the downward continuation of GRAV-D data, and other methods for second check.

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
Fig. 8
Fig. 9
Fig. 10
Fig. 11
Fig. 12
Fig. 13
Fig. 14
Fig. 15
Fig. 16

Similar content being viewed by others

Data availability

The terrestrial gravity data, the airborne gravity data, and the point file of the GSVS17 GPS/leveling data used in this study were provided by the National Geodetic Survey, for the’1 cm Geoid Experiment.’ The airborne gravity data are freely available on https://www.ngs.noaa.gov/GRAV-D/data_products.shtml.

References

  • Alberts B, Klees R (2004) A comparison of methods for the inversion of airborne gravity data. J Geod 78:55–65

    Article  Google Scholar 

  • Amin H, Sjöberg LE, Bagherbandi M (2019) A global vertical datum defined by the conventional geoid potential and the Earth ellipsoid parameters. J Geod 93:1943–1961

    Article  Google Scholar 

  • Brockwell JP, Davis RA (1991) Time series: theory and methods. Springer Verlag, New York

    Book  Google Scholar 

  • Cai J, Grafarend EW, Schaffrin B (2004) The A-optimal regularization parameter in uniform Tykhonov-Phillips regularization—α-weighted BLE-. In: Sansò F (eds) V Hotine-Marussi Symposium on Mathematical Geodesy. International Association of Geodesy Symposia, vol 127. Springer, Berlin, Heidelberg. Doi: https://doi.org/10.1007/978-3-662-10735-5_41

  • Darbeheshti N, Featherstone WE (2010) A review of non-stationary spatial methods for geodetic least-squares collocation. J Spat Sci 55(2):185–204

    Article  Google Scholar 

  • Eicker J, Schall JK (2013) Regional gravity modelling from spaceborne data: case studies with GOCE. Geophys J Int 196(3):1431–1440. https://doi.org/10.1093/gji/ggt485

    Article  Google Scholar 

  • Foroughi I, Abdolreza S, Novák P, Santos MC (2018) Application of radial basis functions for height datum unification. Geosciences 8(10):369

    Article  Google Scholar 

  • Forsberg R (1984) A study of terrain reductions, density anomalies and geophysical inversion methods in gravity field modelling, Dept. of Geod. Sci. Rep., Rep. 355, Ohio State University

  • Forsberg R (1986) Spectral properties of the gravity field in the Nordic Countries. Boll Geodesia e Sc Aff XLV:361–384

    Google Scholar 

  • Forsberg R (1987) A new covariance model for inertial gravimetry and gradiometry. J Geophys Res 92(B2):1305–1310

    Article  Google Scholar 

  • Forsberg R, Olesen A, Bastos L et al (2000) Airborne geoid determination. Earth Planet Sp 52:863–866

    Article  Google Scholar 

  • Forsberg R, Tscherning CC (2014) An overview manual of the GRAVSOFT geodetic gravity field modelling programs, https://ftp.space.dtu.dk/pub/RF/gravsoft_manual2014.pdf

  • Forsberg R, Kaminskis J, Solheim D(1996) Geoid of the Nordic and Baltic Region from gravimetry and satellite altimetry. In: Segawa J, Fujimoto H, Okubo S (ed) Gravity geoid and marine geodesy. IAG symposium series 117: 540–547, Springer Verlag

  • Forsberg R, Olesen A, Keller K (1999) Airborne gravity survey of the North Greenland shelf 1998, Technical Report no. 10, pp. 34, Kort og Matrikelstyrelsen, Copenhagen

  • Forsberg R (2002) Downward continuation of airborne gravity data. In: The 3rd meeting of the international gravity and geoid commission ‘gravity and geoid 2002’. Thessaloniki, Greece

  • Forsberg R (2003) Downward continuation of airborne gravity data. Gravity and Geoid 2002. In: Tziavos (ed) 3rd meeting of the international gravity and geoid commission (IGGC), pp. 51–56, Thessaloniki, Greece August 26–30, 2002, ZITI Editions, Thessaloniki, Greece, ISBN: 960-431-852-7

  • GRAV-D Team (2017a) GRAV-D general airborne gravity data user manual. In: Theresa D, Monica Y, Jeffery J (ed.) Version 2.1. Available DATE. Online at: http://www.ngs.noaa.gov/GRAV-D/data_MS05.shtml

  • GRAV-D Team (2017b). Block MS05 (Mountain South 05); GRAV-D airborne gravity data user manual. In: Monica A. Youngman, Jeffery A Johnson (ed). BETA Available DATE. Online at: http://www.ngs.noaa.gov/GRAV-D/data_MS05.shtml

  • Haagmans RHN, van Gelderen M (1991) Error variances–covariances of GEM-TI: their characteristics and implications in geoid computation. J Geophys Res 96(B12):20011–20022

    Article  Google Scholar 

  • Heiskanen WA, Moritz H (1967) Physical geodesy. Bull Geodesique 86:491–492

    Article  Google Scholar 

  • Holmes SA (2016) Using spherical harmonic expansions for geopotential modeling with airborne gravity, 2016 Airborne gravimetry for geodesy summer school, May 23–27, 2016, in Silver Spring, Maryland, https://www.ngs.noaa.gov/GRAV-D/2016SummerSchool/

  • Holschneider M, Iglewska-Nowak I (2007) Poisson wavelets on the sphere. J Fourier Anal Appl 13:405–419

    Article  Google Scholar 

  • Huang J, Véronneau M, Crowley JW (2019) Experimental determination of geoid models and geopotentials in the US rocky mountains and interior plains: Canadian geodetic survey’s results, 27th IUGG General Assembly, July 8–18, 2019, Montréal, Québec, Canada

  • Huang J, Véronneau M (2005) Applications of downward-continuation in gravimetric geoid modeling: case studies in Western Canada. J Geod 79:135–145

    Article  Google Scholar 

  • Huang J, Véronneau M (2013) Canadian gravimetric geoid model 2010. J Geod 87:771–790. https://doi.org/10.1007/s00190-013-0645-0

    Article  Google Scholar 

  • Huang J, Véronneau M (2015) Assessments of recent GRACE and GOCE release 5 global geopotential models in Canada. Newton’s Bull 5:127–148

    Google Scholar 

  • Huang J (2002) Computational methods for the discrete downward continuation of the earth gravity and effects of lateral topographical mass density variation on gravity and geoid. PhD Thesis, department of geodesy and geomatics engineering, The University of New Brunswick, Fredericton, Canada

  • Hwang C, Hsiao YS, Shih HC, Yang M, Chen KH, Forsberg R, Olesen AV (2007) Geodetic and geophysical results from a Taiwan airborne gravity survey: data reduction and accuracy assessment. J Geophys Res 112:B04407. https://doi.org/10.1029/2005JB004220

    Article  Google Scholar 

  • Jekeli C (1988) The exact transformation between ellipsoidal and spherical harmonic expansions. Manuscr Geod 13:106–113

    Google Scholar 

  • Jekeli C (2010) Correlation modeling of the gravity field in classical geodesy. In: Freeden W, Nashed MZ, Sonar T (eds) Handbook of geomathematics. Springer, Berlin Heidelberg. https://doi.org/10.1007/978-3-642-01546-5_28

    Chapter  Google Scholar 

  • Jekeli C (2016) Theoretical fundamentals of airborne gravimetry, Part II. 2016 airborne gravimetry for geodesy summer school. Silver Spring, Maryland. https://geodesy.noaa.gov/GRAV-D/2016SummerSchool/presentations/day-1/Jekeli_May23_Part2.shtml

  • Jekeli C (1981a) The downward continuation to the Earth’s surface of truncated spherical and sllipsoidal harmonic series of the gravity and height anomalies, OSU Rep. 323, Dept. of Geodetic Science and Surveying, The Ohio State Univ., Columbus

  • Jekeli C (1981b) Alternative methods to smooth the Earth’s gravity field, OSU Rep. 327, Dept. of Geodetic Science and Surveying, The Ohio State Univ., Columbus

  • Jiang T, Li J, Wang Z et al (2011) Solution of ill-posed problem in downward continuation of airborne data. Acta Geod Cartogr Sin 40(6):684–689

    Google Scholar 

  • Kaula WM (1959) Statistical and harmonic analysis of gravity. J Geoph Res 64:2401

    Article  Google Scholar 

  • Kern M (2003) An analysis of the combination and downward continuation of satellite, airborne and terrestrial gravity data. University of Calgary, Calgary, AB. doi:https://doi.org/10.11575/PRISM/22980

  • Kingdon R, Vaníček P (2011) Poisson downward continuation solution by the Jacobi method. J Geod Sci 1:74–81. https://doi.org/10.2478/v10156-010-0009-0

    Article  Google Scholar 

  • Klees R, Tenzer R, Prutkin I, Wittwer T (2008) A data-driven approach to local gravity field modeling using spherical radial basis functions. J Geod 82:457–471

    Article  Google Scholar 

  • Klees R, Slobbe DC, Farahani HH (2018) A methodology for least-squares local quasi-geoid modelling using a noisy satellite-only gravity field model. J Geod 92:431–442. https://doi.org/10.1007/s00190-017-1076-0

    Article  Google Scholar 

  • Klees R, Slobbe DC, Farahani HH (2019) How to deal with the high condition number of the noise covariance matrix of gravity field functionals synthesised from a satellite-only global gravity field model? J Geod 93:29–44. https://doi.org/10.1007/s00190-018-1136-0

    Article  Google Scholar 

  • Krarup T (1969) A contribution to the mathematical foundation of physical geodesy. Geodetisk Institut, Copenhagen

    Google Scholar 

  • Kusche J, Klees R (2002) Regularization of gravity field estimation from satellite gravity gradients. J Geod 76:359–368

    Article  Google Scholar 

  • Li X (2009) Comparing the Kalman filter with a Monte Carlo-based artificial neural network in the INS/GPS vector gravimetric system. J Geod 83:797–804

    Article  Google Scholar 

  • Li X (2011) Strapdown INS/DGPS airborne gravimetry tests in the Gulf of Mexico. J Geod 85:597–605

    Article  Google Scholar 

  • Li X (2018a) Using radial basis functions in airborne gravimetry for local geoid improvement. J Geod 92:471–485

    Article  Google Scholar 

  • Li X (2018b) Modeling the North American vertical datum of 1988 errors in the conterminous United States. J Geod Sci 8(1):1–13

    Article  Google Scholar 

  • Li X, Crowley JW, Holmes SA, Wang YM (2016) The contribution of the GRAV-D airborne gravity to geoid determination in the Great Lakes region. Geophys Res Lett 43:4358–4365

    Article  Google Scholar 

  • Lieb V, Schmidt M, Dettmering D, Börger K (2016) Combination of various observation techniques for regional modeling of the gravity field. J Geophys Res 121(5):3825–3845

    Article  Google Scholar 

  • Lin M, Denker H, Müller J (2019) A comparison of fixed- and free-positioned point mass methods for regional gravity field modeling. J Geodyn 125:32–47. https://doi.org/10.1016/j.jog.2019.01.001

    Article  Google Scholar 

  • Liu M, Huang M, Ouyang Y et al (2016) Test and analysis of downward continuation models for airborne gravity data with regard to the effect of topographic height. Acta Geod Cartogr Sin 45(5):521–530

    Google Scholar 

  • Liu Q, Schmidt M, Sánchez L et al (2020) Regional gravity field refinement for (quasi-) geoid determination based on spherical radial basis functions in Colorado. J Geod 94:99

    Article  Google Scholar 

  • Martinec Z (1996) Stability investigations of a discrete downward continuation problem for geoid determination in the Canadian rocky mountains. J Geod 70:805–828

    Article  Google Scholar 

  • Milbert DG (1999) The dilemma of downward continuation. AGU 1999spring meeting, 1–4 June, 1999, Boston

  • Moritz H (1980) Advanced physical geodesy. Wichmann Verlag, Karlsruhe, p 499p

    Google Scholar 

  • Moritz H (1972) Advanced least-squares methods, Dep. of Geodetic Science and Surveying, Ohio State Univ. Rep. No. 175

  • Naeimi M (2013) Inversion of satellite gravity data using spherical radial base functions, PhD thesis, Munchen

  • Novák P, Heck B (2002) Downward continuation and geoid determination based on band-limited airborne gravity data. J Geod 76:269–278

    Article  Google Scholar 

  • Novák P, Kern M, Schwarz KP, Sideris MG, Heck B, Ferguson S, Hammada Y, Wei M (2003) On geoid determination from airborne gravity. J. Geod 76:510–522

    Article  Google Scholar 

  • Pail R, Reguzzoni M, Sansò F, Kühtreiber N (2010) On the combination of global and local data in collocation theory. Stud Geophys Geod 54(2):195–218

    Article  Google Scholar 

  • Pavlis NK, Holmes SA, Kenyon SC, Factor JK (2012) The development and evaluation of the earth gravitational model 2008 (EGM2008). J Geophys Res 117:B04406. https://doi.org/10.1029/2011JB008916

    Article  Google Scholar 

  • Pavlis NK (1998) The block-diagonal least-squares approach, in Development and preliminary investigation. In: The development of the joint NASA GSFC and the national imagery and mapping. Agency (NIMA) Geopotential Model EGM96, NASA Tech. Publ., TP-1998-206861, sect. 8.2.2, pp. 8-4–8-5, NASA Goddard Space Flight Cent., Washington, D. C

  • Rexer M, Hirt C, Claessens S, Tenzer R (2016) Layer-based modelling of the earth’s gravitational potential up to 10-km scale in spherical harmonics in spherical and ellipsoidal approximation. Surv Geophys 37(6):1035–1074. https://doi.org/10.1007/s10712-016-9382-2

    Article  Google Scholar 

  • Rummel R, Schwarz KP, Gerstl M (1979) Least square collocation and regularization. Belletin Geodesique 53:343–361

    Article  Google Scholar 

  • Sánchez L, Ågren J, Huang J et al (2021) Strategy for the realisation of the international height reference system (IHRS). J Geod 95:33. https://doi.org/10.1007/s00190-021-01481-0

    Article  Google Scholar 

  • Schaffrin B (2006) A note on constrained total least-squares estimation. Linear Algebra Appl 417(1):245–258

    Article  Google Scholar 

  • Schmidt M, Fengler M, Mayer-Guerr T, Eicker A, Kusche J, Sanchez L, Han SC (2007) Regional gravity field modeling in terms of spherical base functions. J Geod 81(1):17–38. https://doi.org/10.1007/s00190-006-0101-5

    Article  Google Scholar 

  • Schwarz KP (1978) Geodetic improperly posed problems and their regularization, Lecture Notes of the Second Int. School of Advance Geodesy, Erice

  • Sjöberg LE (2007) The topographic bias by analytical continuation in physical geodesy. J Geod 81:345–350. https://doi.org/10.1007/s00190-006-0112-2

    Article  Google Scholar 

  • Slobbe DC, Klees R, Farahani HH, Huisman L, Alberts B, Voet P, de Doncker F (2019) The impact of noise in a GRACE/GOCE global gravity model on a local quasi-geoid. JGR Solid Earth 124(3):3219–3237. https://doi.org/10.1029/2018JB016470

    Article  Google Scholar 

  • Slobbe DC (2013) Roadmap to a mutually consistent set of offshore vertical reference frames. PhD thesis, Delft University of Technology

  • Smith DA, Holmes SA, Li X et al (2013) Confirming regional 1 cm differential geoid accuracy from airborne gravimetry: the geoid slope validation survey of 2011. J Geod 87:885–907

    Article  Google Scholar 

  • Tscherning CC (2013) Geoid determination by 3D least-squares collocation. In: Sansò F, Sideris M (eds) Geoid determination. Lecture notes in earth system sciences, vol 110. Springer, Berlin Heidelberg

    Google Scholar 

  • Tscherning CC, Rapp RH (1974) Closed covariance expressions for gravity anomalies, geoid undulations, and deflections of the vertical implied by anomaly degree variance models. Dep. of Geodetic Science and Surveying, Ohio State Univ. Rep. no. 208

  • van Westrum D, Ahlgren K, Hirt C, Guillaume S (2021) A geoid slope validation survey (2017) in the rugged terrain of Colorado, USA. J Geod 95:9. https://doi.org/10.1007/s00190-020-01463-8

    Article  Google Scholar 

  • Vaníček P, Sun W, Ong P (1996) Downward continuation of Helmert’s gravity. J Geod 71:21–34

    Article  Google Scholar 

  • Véronneau M, Huang J (2016) The Canadian geodetic vertical datum of 2013 (CGVD2013). Geomatica 70(1):9–19

    Article  Google Scholar 

  • Wang YM, Sánchez L, Ågren J et al (2021) Colorado geoid computation experiment: overview and summary. J Geod 95:127. https://doi.org/10.1007/s00190-021-01567-9

    Article  Google Scholar 

  • Willberg M, Zingerle P, Pail R (2019) Residual least-squares collocation: use of covariance matrices from high-resolution global geopotential models. J Geod 93:1739–1757

    Article  Google Scholar 

  • Willberg M, Zingerle P, Pail R (2020) Integration of airborne gravimetry data filtering into residual least-squares collocation: example from the 1 cm geoid experiment. J Geod 94:75

    Article  Google Scholar 

  • Xu P (1992) Determination of surface gravity anomalies using gradiometric observables. Geophys J Int 110:321–332

    Article  Google Scholar 

  • Xu P, Rummel R (1994) Generalized ridge regression with applications in determination of potential fields. Manuscr Geodaet 20:8–20

    Google Scholar 

  • Zhao Q, Xu X, Forsberg R, Strykowski G (2018) Improvement of downward continuation values of airborne gravity data in Taiwan. Remote Sens 10:1951

    Article  Google Scholar 

Download references

Acknowledgements

The authors thank the comments from the three reviewers, and the Associate Editor, Prof. Tziavos. Many appreciations are also given to NGS former Chief Scientist, Dr. Milbert and Mr. Véronneau at NRCan for the useful discussions, comments and suggestions during the internal reviewing process. The first author also thanks the support of this study from Dr. Roman, Chief Geodesist at NGS.

Author information

Authors and Affiliations

Authors

Contributions

Conception or design of the work was contributed by XL, JH, RK, RF, MW, CS, RP, and CH; Data analysis was done by XL; Geoid modeling was done by JH; Writing—original draft preparation was contributed by XL and JH; Writing—review and editing were done by RK, RF, MW, and CH; Critical revisions of the article were done by RK and JH.

Corresponding author

Correspondence to X. Li.

Appendices

Appendix A

1.1 The derivation of the equations of downward continuation of airborne gravity data using various approaches

1.1.1 Spherical harmonic analysis (SHA)

In spherical approximation, band-limited (full/residual) gravity disturbances from airborne gravimetry can be modelled as

$$ \begin{aligned} & \delta g\left( {r~,\phi ~,\lambda ~} \right) \approx - \frac{{\partial T}}{{\partial r}} \\ & \quad = \frac{{{\text{GM}}}}{{R^{2} }}\mathop \sum \limits_{{n = n1}}^{{n2}} \left( {\frac{R}{{r~}}} \right)^{{n + 2}} \left( {n + 1} \right)\\ &\quad \times \mathop \sum \limits_{{m = 0}}^{n} \left( {\bar{c}_{{{\text{nm}}}} ~\cos \left( {m\lambda } \right) + \bar{s}_{{{\text{nm}}}} ~\sin \left( {m\lambda } \right)} \right)\bar{p}_{{{\text{nm}}}} ~\left( {\sin \phi } \right) \\ \end{aligned} $$
(13)

where \(T\) is the gravity disturbing potential; \({\text{GM}}\) is the product of the Universal gravitational constant and the assigned reference mass; \(R\) is the mean Earth’s radius; \(\overline{p}_{{{\text{nm}}}}\) are the fully normalized associated Legendre functions of degree \(n\) and order \(m\); \(\left\{ {\overline{c}_{{{\text{nm}}}} ,\overline{s}_{{{\text{nm}}}} } \right\}\) are the Stokes’s coefficients to be solved from the airborne gravity disturbances \({\updelta }g\); and the limits of \(n1\) to \(n2\) define the bandwidth of the airborne gravity disturbances.

However, solving for all coefficients \(\left\{ {\overline{c}_{{{\text{nm}}}} ,\overline{s}_{{{\text{nm}}}} } \right\}\) from airborne data given inside a local area is not possible; in fact, there are infinitely many solutions to this problem. To foster one particular solution, one may assume zero gravity data outside the local area.

This treatment will definitely reduce the power in the spherical harmonic coefficients because of applying the global function for a representation of space-limited data. This set of coefficients are no longer representing the true gravity field of the Earth, especially outside the study region (Jekeli 2016). An intuitive explanation is that all the data of the world implies that \(\left\{ {\overline{c}_{{{\text{nm}}}} ,\overline{s}_{{{\text{nm}}}} } \right\}\) are zeros, except only a small amount of regional data says they are not zeros. Note that the spherical harmonic functions, \(\overline{p}_{{{\text{nm}}}} \left( {\sin \phi } \right)\cos \left( {m\lambda } \right) \) and \(\overline{p}_{{{\text{nm}}}} \left( {\sin \phi } \right)\sin \left( {m\lambda } \right)\), are global functions; using regional data, it is impossible to estimate all spherical harmonic coefficients.

In addition to the theoretical shortcomings of using SHA to downward continue airborne data, numerically, Eq. (13) cannot be established directly on the original airborne survey point \( i\left( {r_{i} ,\phi_{i} ,\lambda_{i} } \right)\) in order to utilize the EGM208 block diagonal engine to quickly solve the 4 –million spherical harmonic coefficients. In practice, the point airborne data has to be reduced from the flight altitude onto a grid on the surface of a larger (in terms of semi-major axis) reference ellipsoid that is located in the mean flight height in order to solve the spheroidal harmonic coefficients quickly by using the well-known block diagonal method (Pavlis 1998), which are then transformed into spherical harmonics by using Jekeli’s method (1988). Finally, the spherical harmonic coefficients are scaled back to the desired system. The details of spheroidal harmonic analysis and their transformations to the spherical harmonic are well-known and will not be repeated here.

In the GRAV-D airborne gravity data, the flight heights can vary by several thousand meters in a typical survey campaign. The real flights are not on a mean flight height as assumed by some algorithms. A novel iterative scheme (Smith et al. 2013) described by the following steps was developed to downward continue the real data from varying flight heights on to an inflated ellipsoid whose mean radius is \(R^{{{\text{big}}}}\), and its surface is located at the numerical mean flight height. Then the spherical harmonic coefficients are scaled back by using the ratio between \(R^{{{\text{big}}}}\) and \(R_{{}}^{{}}\).

  1. (1)

    First, let’s assume the residual airborne gravity disturbances are given on the surface of an inflated ellipsoid and solve the spherical harmonic coefficients:

    $$ \begin{aligned}& \left\{ {\overline{c}_{{{\text{nm}}}}^{T0} ,\overline{s}_{{{\text{nm}}}}^{T0} } \right\}\\ &\quad= {\mathcal{F}}_{0}^{ - 1} \left\{ {{\updelta }\underline {g}_{0} \left( {r_{i} - h,\phi_{i} ,\lambda_{i} } \right) = {\updelta }\underline {g} \left( {r_{i} ,\phi_{i} ,\lambda_{i} } \right)} \right\}, \end{aligned}$$
    (14)

    where \({\updelta }\underline {g}_{0} \left( {r_{i} - h_{i} ,\phi_{i} ,\lambda_{i} } \right)\) are the pseudo gravity disturbances on the reference ellipsoid; \({\mathcal{F}}_{0}^{ - 1} \left\{ \cdot \right\}\) represent spherical harmonic analysis operator for the block diagonal method (for simplicity of description the steps to transform the ellipsoid harmonics into spherical harmonics are neglected, but it is applied in the computation);

  2. (2)

    Then, the following two terms are generated for the next iteration.

The first term is called the misclosure:

$$ {\updelta }\underline {g}^{M1} = {\mathcal{F}}_{0}^{{\left( {r_{i} ,\phi_{i} ,\lambda_{i} } \right)}} \left\{ {\overline{c}_{{{\text{nm}}}}^{T0} ,\overline{s}_{nm}^{T0} } \right\} - {\updelta }\underline {g} , $$
(15)

where \({\mathcal{F}}_{0}^{{}} \left\{ \cdot \right\}\) is the spherical harmonic synthesis operator (SHS) at the original airborne survey points at the real altitudes based on \(\left\{ {\overline{c}_{{{\text{nm}}}}^{T0} ,\overline{s}_{{{\text{nm}}}}^{T0} } \right\}\). Basically, this term shows the errors in the spherical harmonic coefficients due to the assumption made in step (1).

The second term is the DC term:

$$ {\updelta }\underline {g}^{{{\text{DC1}}}} = {\mathcal{F}}_{0}^{{\left( {r_{i} ,\phi_{i} ,\lambda_{i} } \right)}} \left\{ {\overline{c}_{{{\text{nm}}}}^{T0} ,\overline{s}_{{{\text{nm}}}}^{T0} } \right\} - {\mathcal{F}}_{0}^{{\left( {r_{i} - h_{i} ,\phi_{i} ,\lambda_{i} } \right)}} \left\{ {\overline{c}_{{{\text{nm}}}}^{\delta 0} ,\overline{s}_{{{\text{nm}}}}^{\delta 0} } \right\} $$
(16)
  1. (3)

    Based on the DC term, a new observation as shown in Eq. (17) is assumed to be given on the surface of the bigger ellipsoidal surface for the next round of SHA as shown in Eq. (18).

    $$ {\updelta }\underline {g}_{1} \left( {r_{i} - h_{i} ,\phi_{i} ,\lambda_{i} } \right) = {\updelta }\underline {g} \left( {r_{i} ,\phi_{i} ,\lambda_{i} } \right) - {\updelta }\underline {g}^{{{\text{DC1}}}} , $$
    (17)
    $$ \left\{ {\overline{c}_{{{\text{nm}}}}^{T1} ,\overline{s}_{{{\text{nm}}}}^{T1} } \right\} = {\mathcal{F}}_{1}^{ - 1} \left\{ {{\updelta }\underline {g}_{1} \left( {r_{i} - h,\phi_{i} ,\lambda_{i} } \right)} \right\}, $$
    (18)

Another misclosure term is obtained by:

$$ {\updelta }\underline {g}^{M2} = {\mathcal{F}}_{1}^{{\left( {r_{i} ,\phi_{i} ,\lambda_{i} } \right)}} \left\{ {\overline{c}_{{{\text{nm}}}}^{T1} ,\overline{s}_{{{\text{nm}}}}^{T1} } \right\} - {\updelta }\underline {g} , $$
(19)
  1. (4)

    In this step, the DC term of the misclosure term is computed according to:

    $$ \left\{ {\overline{c}_{{{\text{nm}}}}^{T2} ,\overline{s}_{{{\text{nm}}}}^{T2} } \right\} = {\mathcal{F}}_{2}^{ - 1} \left\{ {{\updelta }\underline {g}^{M2} } \right\}, $$
    (20)
    $$\begin{aligned} {\updelta }\underline{\Delta g}^{{{\text{DC2}}}}& = {\mathcal{F}}_{2}^{{\left( {r_{i} ,\phi_{i} ,\lambda_{i} } \right)}} \left\{ {\overline{c}_{{{\text{nm}}}}^{T2} ,\overline{s}_{{{\text{nm}}}}^{T2} } \right\} \\& \quad - {\mathcal{F}}_{2}^{{\left( {r_{i} - h_{i} ,\phi_{i} ,\lambda_{i} } \right)}} \left\{ {\overline{c}_{{{\text{nm}}}}^{T2} ,\overline{s}_{{{\text{nm}}}}^{T2} } \right\}, \end{aligned}$$
    (21)
  2. (5)

    Finally, the original airborne gravity is reduced from the flight trajectories onto the surface of the inflated ellipsoid by:

    $$ {\updelta }\underline {g}_{3} \left( {r_{i} - h_{i} ,\phi_{i} ,\lambda_{i} } \right) = {\updelta }\underline {g}_{1} - \left( {{\updelta }\underline {g}^{M2} - {\updelta }\underline {g}^{{{\text{DC2}}}} } \right), $$
    (22)

Finally, the spherical harmonic coefficients for the original airborne data are given as

$$ \left\{ {\overline{c}_{{{\text{nm}}}}^{T} ,\overline{s}_{{{\text{nm}}}}^{T} } \right\} = {\mathcal{F}}_{3}^{ - 1} \left\{ {{\updelta }\underline {g}_{3} } \right\}, $$
(23)

Note that a gridding step is still needed to interpolate the scattered airborne gravity data onto the nodes of a grid (one data grid and one DEM grid) to start the iterations described above.

1.1.2 Least squares collocation (LSC) method

Forsberg’s (1987) covariance model is based on a power spectral model for a planar residual gravity field, matching the Kaula rule up in the main band frequency band up to a “depth” parameter D (equivalent to the use of a Bjerhammar sphere, but twice the magnitude). This gives rise to a gravity covariance function \({C}_{\delta g,\delta g}\) of shape − log(z + r), where z = z1 + z2 + D is a height parameter, and r the slant range between the two covariance points. To avoid singularities in the geoid terms for a fully self-consistent model, several log-terms are needed to be combined through a deeper “compensation depth” T, giving a relatively complicated expression of form

$$ \begin{aligned} & C_{{\delta g,\delta g}} \left\{ {P\left( {\phi _{i} ,\lambda _{i} ,h_{i} } \right),Q\left( {\phi _{j} ,\lambda _{j} ,h_{j} } \right)} \right\}\\ &\quad = \frac{{\sigma ^{2} }}{{3\log \left( {D + T} \right) - 3\log \left( {D + 2T} \right) + \log \left( {D + 3T} \right) - \log \left( D \right)}}\\ &\qquad \times \left\{ \begin{gathered} 3\log \left( {D + T + h_{i} + h_{j} + \sqrt {s + \left( {D + T + h_{i} + h_{j} } \right)^{2} } } \right) \hfill \\ - 3\log \left( {D + 2T + h_{i} + h_{j} + \sqrt {s + \left( {D + 2T + h_{i} + h_{j} } \right)^{2} } } \right) \hfill \\ + \log \left( {D + 3T + h_{i} + h_{j} + \sqrt {s + \left( {D + 3T + h_{i} + h_{j} } \right)^{2} } } \right) \hfill \\ - \log \left( {D + h_{i} + h_{j} + \sqrt {s + \left( {D + h_{i} + h_{j} } \right)^{2} } } \right) \hfill \\ \end{gathered} \right\} \end{aligned} $$
(24)

where \(\sigma_{{}}^{2}\) is the data variance; \(D\) is the depth to the high frequency attenuation layer; \(T\) is the thickness to bottom layer, low frequency attenuation; and \(s = \sqrt {x_{{}}^{2} + y_{{}}^{2} }\) the planar distance (approximated here with \(x = 111.11{\text{km}} \times \left( {\lambda_{i} - \lambda_{j} } \right)\cos \overline{\phi }\) and \(y = 111.11{\text{km}} \times \left( {\phi_{i} - \phi_{j} } \right)\cos \overline{\phi }\); \(h_{i}^{{}}\) and \(h_{j}^{{}}\) are the flight height in km).

The covariance between gravity disturbance, \(\delta g,\) and height anomaly, \({\upzeta }\), based on the covariance model of gravity anomalies is (Forsberg, 1987):

$$ \begin{gathered} C_{\delta g,\zeta } \left\{ {P\left( {\phi_{i} ,\lambda_{i} ,h_{i} } \right),Q\left( {\phi_{j} ,\lambda_{j} ,h_{j} } \right)} \right\} = - \frac{1}{\gamma }\frac{{\sigma^{2} }}{{3\log \left( {D + T} \right) - 3\log \left( {D + 2T} \right) + \log \left( {D + 3T} \right) - \log \left( D \right)}} \hfill \\ \quad \times\left\{ \begin{gathered} \left( {\sqrt {s + \left( {D + h_{i} + h_{j} } \right)^{2} } - \left( {D + h_{i} + h_{j}^{{}} } \right)\log \left( {D + h_{i} + h_{j} + \sqrt {s + \left( {D + h_{i} + h_{j} } \right)^{2} } } \right)} \right) \hfill \\ - 3\left( {\sqrt {s + \left( {D + T + h_{i} + h_{j} } \right)^{2} } - \left( {D + T + h_{i} + h_{j} } \right)\log \left( {D + T + h_{i} + h_{j} + \sqrt {s + \left( {D + T + h_{i} + h_{j} } \right)^{2} } } \right)} \right) \hfill \\ + 3\left( {\sqrt {s + \left( {D + 2T + h_{i} + h_{j} } \right)^{2} } - \left( {D + 2T + h_{i} + h_{j} } \right)\log \left( {D + 2T + h_{i} + h_{j} + \sqrt {s + \left( {D + 2T + h_{i} + h_{j} } \right)^{2} } } \right)} \right) \hfill \\ - \left( {\sqrt {s + \left( {D + 3T + h_{i} + h_{j} } \right)^{2} } - \left( {D + 3T + h_{i} + h_{j} } \right)\log \left( {D + 3T + h_{i} + h_{j} + \sqrt {s + \left( {D + 3T + h_{i} + h_{j} } \right)^{2} } } \right)} \right) \hfill \\ \end{gathered} \right\} \hfill \\ \end{gathered} $$
(25)

Given the covariance function, the prediction of gravity anomalies and height anomalies including error variances using LSC can be implemented by the well-known collocation equations:

$$ \widehat{\delta g}^{{{\text{pred}}}} = C_{{\delta g^{{{\text{pred}}}} ,\delta g^{{{\text{obs}}}} }} \left( {C_{{\delta g^{{{\text{obs}}}} ,\delta g^{{{\text{obs}}}} }} + D_{n,n} } \right)^{ - 1} \left[ {\begin{array}{*{20}c} {\delta g_{1}^{{{\text{obs}}}} } \\ {\begin{array}{*{20}c} {\delta g_{2}^{{{\text{obs}}}} } \\ \vdots \\ \vdots \\ \end{array} } \\ {\delta g_{n}^{{{\text{obs}}}} } \\ \end{array} } \right], $$
(26)
$$ \begin{aligned} D\{ \widehat{\delta g}^{{{\text{pred}}}} \} &= C_{{\delta g,\underline{\delta g} }} - C_{{\delta g^{{{\text{pred}}}} ,\delta g^{{{\text{obs}}}} }} \left( {C_{{\delta g^{{{\text{obs}}}} ,\delta g^{{{\text{obs}}}} }} + D_{n,n} } \right)^{ - 1}\\ & \quad C_{{\delta g^{{{\text{pred}}}} ,\underline{\delta g}^{{{\text{obs}}}} }}^{T} ,\end{aligned} $$
(27)
$$ {\hat{\zeta }}^{{{\text{pred}}}} = C_{{{\upzeta }^{{{\text{pred}}}} ,\underline{\delta g}^{{{\text{obs}}}} }} \left( {C_{{\delta g^{{{\text{obs}}}} ,\delta g^{{{\text{obs}}}} }} + D_{n,n} } \right)^{ - 1} \left[ {\begin{array}{*{20}c} {\delta g_{1}^{{{\text{obs}}}} } \\ {\begin{array}{*{20}c} {\delta g_{2}^{{{\text{obs}}}} } \\ \vdots \\ \vdots \\ \end{array} } \\ {\delta g_{n}^{{{\text{obs}}}} } \\ \end{array} } \right], $$
(28)
$$ \begin{aligned} D\{ {\hat{\zeta }}^{{{\text{pred}}}} \} &= C_{{{\upzeta },{\upzeta }}} - C_{{{\upzeta }^{{{\text{pred}}}} ,\underline{\delta g}^{{{\text{obs}}}} }} \left( {C_{{\underline{\delta g}^{{{\text{obs}}}} ,\underline{\delta g}^{{{\text{obs}}}} }} + D_{n,n} } \right)^{ - 1} \\& \quad C_{{{\upzeta }_{{}}^{{{\text{pred}}}} ,\underline{\delta g}^{{{\text{obs}}}} }}^{T} ,\end{aligned} $$
(29)

where \(C_{{{\upzeta },{\upzeta }}} = \frac{1}{{\gamma^{2} }}\left[ {\frac{3}{4}zr + \left( {\frac{1}{4}r^{2} - \frac{3}{4}z^{2} } \right)\log \left( {z + r} \right)} \right].\)

In the planar approximation, the covariance model is described only by three parameters, i.e., gravity variance \(\sigma_{{}}^{2}\), high-frequency attenuation depth parameter \(D\), and low-frequency attenuation depth parameter \(T\). Due to its simplicity, it is widely used in processing airborne gravity data (Forsberg et al. 1996, 1999, 2000, 2003). However, the simplified covariance function cannot always yield a good fit of the data in the remove-compute-restore scenario, especially for the geoid part, as the covariance functions have difficulties to approximate the negative covariances coming from removing a reference field (which again is a consequence of the planar approximation).

1.1.3 A.3 Residual LSC

Residual least-squares collocation is a modified form of LSC (Willberg et al. 2019) in the context of remove-compute-restore (RCR). In RLSC, the covariance modeling is taken from the global gravity field modeling (GGM) and forward topographic modeling instead of empirical parameterization. The general form of RLSC reads as (Willberg et al. 2019):

$$ {\mathbf{s}} = \left( {{\mathbf{C}}_{{\hat{s}\hat{l}}}^{m} + {\mathbf{C}}_{{\hat{s}\hat{l}}}^{t} } \right)\left( {{\mathbf{C}}_{ll} + {\mathbf{C}}_{{\hat{l}\hat{l}}}^{m} + {\mathbf{C}}_{{\hat{l}\hat{l}}}^{t} } \right)^{ - 1} \left( {{\mathbf{l}} - {\hat{\mathbf{l}}}^{m} - {\hat{\mathbf{l}}}^{t} } \right) + {\hat{\mathbf{s}}}^{m} + {\hat{\mathbf{s}}}^{t} , $$
(30)

where \({\mathbf{s}}\) is the signal, \({\mathbf{l}}\) is the observable, and the super-scripts \(m\) and \(t\) are representing information from GGM and topography, respectively. The circumflexes indicates the corresponding quantities are a-priori estimates from GGM and topographic modeling. \({\hat{\mathbf{l}}}^{m} , {\hat{\mathbf{l}}}^{t}\) are linked to the remove step, \({\hat{\mathbf{s}}}^{m} ,{\hat{\mathbf{s}}}^{t}\) are linked to the restore step.

From Eq. (30), we can see that RLSC uses separated error covariance matrices \({\mathbf{C}}\) to describe signal correlations from every input source. In particular, \({\mathbf{C}}_{{\widehat{ \cdot }\widehat{ \cdot }}}^{m}\) includes the full covariance information from a high-resolution GGM, which is not included in LSC. Accordingly, it considers accuracies and correlations of the high-resolution GGM and models the gravity field of the Earth as anisotropic and location-dependent (up to the maximum degree of the GGM). This more realistic representation of the gravity field should generally improve the quality of a regional gravity field model.

A more detailed differentiation between LSC and RLSC is explained in Willberg et al. (2019), which also highlights corresponding similarities to specific realizations of LSC that already include accuracy information from a GGM in different forms (e.g., Haagmans & van Gelderen, 1991; Pail et al. 2010).

1.1.4 Radial basis function (RBF)

The (residual) gravity disturbing potential is written as

$$ \begin{aligned}\delta T\left( {r_{p} ,\phi_{p} ,\lambda_{p} } \right)& = \frac{{{\text{GM}}}}{{R_{{}}^{{}} }}\mathop \sum \limits_{q = 1}^{{Q_{L} }} \alpha_{q} \mathop \sum \limits_{n = n1}^{n2} \left( {\frac{R}{{r_{p} }}} \right)^{n + 1}\\ & \quad \times b_{n} \sqrt {2n + 1} \overline{p}_{n} \left( {\frac{{\vec{r}_{p} \cdot \vec{r}_{q} }}{{\left| {\vec{r}_{p} } \right|\left| {\vec{r}_{q} } \right|}}} \right)\end{aligned} $$
(31)

where \(b_{n}\) is a set of parameters that control the type of RBF (Schmidt et al. 2007), which determine the spectrum of the RBF, \(\vec{r}_{q}\) denotes the location of the center of the RBF,\( \overline{p}_{n}\) is the \(4\pi\)-normalized Legendre polynomial of degree n, and \(\alpha_{q}^{{}}\) are the parameters to be estimated from the data, and \(Q_{L}\) is the total number of coefficients. Then, in spherical approximation, the residual gravity disturbance is given by:

$$\begin{aligned} &\delta \delta g\left( {r_{p} ,\phi_{p} ,\lambda_{p} } \right) = - \frac{\partial \delta T}{{\partial r_{p} }} = \frac{{{\text{GM}}}}{{R^{2} }}\mathop \sum \limits_{q = 1}^{{Q_{L} }} \alpha_{q} \\& \quad \times\mathop \sum \limits_{n = n1}^{n2} \left( {\frac{R}{{r_{p} }}} \right)^{n + 2} b_{n} \left( {n + 1} \right)\sqrt[{}]{2n + 1}\overline{p}_{n} \left( {\frac{{\vec{r}_{p} \cdot \vec{r}_{q} }}{{\left| {\vec{r}_{p} } \right|\left| {\vec{r}_{q} } \right|}}} \right),\end{aligned} $$
(32)

Eq. (32) provides the functional model for the estimation of the RBF parameters using ordinary least-squares techniques. Note that the computation of residual gravity disturbances at the Earth’s surface requires an extra synthesis step once the RBF parameters have been computed. Conceptually, using least-squares to estimate the RBF parameters allows the use of the original data along the flight directory without any spatial interpolation prior to the estimation of the RBF coefficients. This convenience avoids all of the complicated data pre-processing needed in steps (1) to (5) in A.1. Furthermore, the spatial concentration of a RBF allows for a parameterization of a regional gravity field requiring much less coefficients compared to the use of spherical harmonics (e.g., Klees et al. 2008; Lieb et al. 2016; Li 2018a).

The performance of the RBF least-squares approach critically depends on the definition of the RBF network, i.e., the location, including the depth (for some types of RBFs), and distribution of the RBFs. There are many ways to find a suitable RBF network (e.g., Schmidt et al. 2007, Klees et al. 2008, Eicker et al. 2013, Naeimi 2013, Lin et al. 2019, Foroghi et al. 2018). In this study, we use the nodes of a Reuter grid to position the RBFs horizontally (Eicker et al. 2013, Naeimi 2013, Lieb et al. 2016, Klees et al. 2018, Klees et al. 2019).

After setting up the system, the observation equations can be written as a Gauss-Markov Model (Schaffrin 2006):

$$ {\mathbf{E}}\left( {\mathbf{y}} \right) = A{\varvec{\xi}}, D\left( y \right) = \sigma_{0}^{2} P^{ - 1} , $$
(33)

where \({\mathbf{y}}\) is a \(m \times 1\) column vector that contains the residual gravity disturbances, \(A\) is a \(m \times k\) design matrix with \(k = Q_{L}^{{}}\) number of basis functions used, and \({\varvec{\xi}}\) is a \(k \times 1\) vector comprising the RBF parameters. The regularized ordinary least-squares solution is given by (Eq. 78 Schmidt et al. 2007):

$$ \hat{\user2{\xi }} = \left( {A^{T} PA + {{\varvec{\upkappa}}}^{{{\mathbf{om}}}} {\mathbf{I}}} \right)^{ - 1} (A^{T} {\mathbf{Py}}), $$
(34)

where the optimal regularization parameter \({{\varvec{\upkappa}}}_{{{\mathbf{om}}}}\) is computed using Algorithm B in (Xu 1992) and

$$ {\varvec{D}}\{ \widehat{{{\varvec{\xi}}\} }} = \sigma_{y}^{2} \left( {A^{T} A + {{\varvec{\upkappa}}}_{{{\mathbf{om}}}} {\mathbf{I}}} \right)^{ - 1} \left( {A^{T} A} \right)\left( {A^{T} A + {{\varvec{\upkappa}}}_{{{\mathbf{om}}}} {\mathbf{I}}} \right)^{ - 1} , $$
(35)

The downward continued gravity disturbance is simply given by:

$$ {\hat{\mathbf{y}}} = A^{dc} \hat{\user2{\xi }};\user2{ }D\left\{ {{\hat{\mathbf{y}}}} \right\} = A^{dc} D\left\{ {\hat{\user2{\xi }}} \right\}A^{dcT} , $$
(36)

1.1.5 Inverse Poisson method

The formula for the upward continuation of a harmonic function \(r_{p} {\updelta }g\left( {r_{p} ,\phi_{p} ,\lambda_{p} } \right)\) at flight altitude P (\(r_{p} ,\phi_{p} ,\lambda_{p}\)) from a spherical surface (\(\sigma_{R}^{{}} \left( {Q_{{}}^{^{\prime}} } \right)\)) is given by the Poisson integral as (e.g., Alberts and Klees 2004):

$$\begin{aligned} &r_{p} {\updelta }g\left( {r_{p} ,\phi_{p} ,\lambda_{p} } \right) = \frac{1}{{4\pi R^{2} }}\iint\limits_{{\sigma_{R}^{{}} }} {R{\updelta }g\left( {r_{q}^{^{\prime}} ,\phi_{q}^{^{\prime}} ,\lambda_{q}^{^{\prime}} } \right)}\\ & \quad {\wp \left( {R,\frac{{\vec{r}_{p} \cdot \vec{r}_{q}^{^{\prime}} }}{{\left| {\vec{r}_{p} } \right|\left| {\vec{r}_{q}^{^{\prime}} } \right|}},r_{p} } \right)d\sigma_{R} \left( {Q^{^{\prime}} } \right),} \end{aligned}$$
(37)

where \(\wp \left( {R,\frac{{\vec{r}_{p} \cdot \vec{r}_{q}^{^{\prime}} }}{{\left| {\vec{r}_{p} } \right|\left| {\vec{r}_{q}^{^{\prime}} } \right|}},r_{p} } \right)\) is the Poisson kernel given by

$$ \begin{aligned}&\wp \left( {R,\frac{{\vec{r}_{p} \cdot \vec{r}_{q}^{^{\prime}} }}{{\left| {\vec{r}_{p} } \right|\left| {\vec{r}_{q}^{^{\prime}} } \right|}},r_{p} } \right) = \mathop \sum \limits_{n = 0}^{\infty } \left( {\frac{R}{{r_{p} }}} \right)^{n + 1} \sqrt {2n + 1} \overline{p}_{n} \left( {\frac{{\vec{r}_{p} \cdot \vec{r}_{q}^{^{\prime}} }}{{\left| {\vec{r}_{p} } \right|\left| {\vec{r}_{q}^{^{\prime}} } \right|}}} \right) \\&\quad = \mathop \sum \limits_{n = 0}^{\infty } 2n + 1\left( {\frac{R}{{r_{p} }}} \right)^{n + 1} P_{n} \left( {\frac{{\vec{r}_{p} \cdot \vec{r}_{q}^{^{\prime}} }}{{\left| {\vec{r}_{p} } \right|\left| {\vec{r}_{q}^{^{\prime}} } \right|}}} \right) = R\frac{{r_{p}^{2} - R^{2} }}{{l^{3} }}, \end{aligned}$$
(38)

The distances between \(P\left( {r_{p} ,\phi_{p} ,\lambda_{p} } \right)\) and \(Q^{\prime}\left( {r_{q}^{^{\prime}} ,\phi_{q}^{^{\prime}} ,\lambda_{q}^{^{\prime}} } \right)\) is

$$ l = \sqrt {r_{p}^{2} + R^{2} - 2rR\frac{{\vec{r}_{p} \cdot \vec{r}_{q}^{^{\prime}} }}{{\left| {\vec{r}_{p} } \right|\left| {\vec{r}_{q}^{^{\prime}} } \right|}}} = \sqrt {r_{p}^{2} + R^{2} - 2\vec{r}_{p}^{{}} \cdot \vec{r}_{q}^{^{\prime}} } $$
(39)

Thus, the relationship between the observed gravity disturbances at altitude and the counterparts on the geoid in a spherical approximation is given by:

$${{\updelta }}g\left( P \right) = \frac{{R^{2} }}{{4\pi r_{p} }}\iint\limits_{{\sigma _{R} }} {{{\updelta }}g\left( {Q^{'} } \right)\frac{{r_{p}^{2} - R^{2} }}{{l^{3} }}{\rm{d\sigma }}_{R} \left( {Q^{'} } \right)}$$
(40)

Solving \(\updelta g\left({Q}^{^{\prime}}\right)\) from \(\updelta g\left(P\right)\) is called the DC of the gravity disturbance or the inverse Poisson problem. In general, it belongs to the Fredholm integral equation of the first kind in mathematics, which is ill-posed due to the amplification of noise when dealing with a high-resolution field. It is essential to stabilize the Poisson DC using regularization or smoothing (see e. g. Martinec 1996).

For the DC of airborne data, the inversion is done for a regional area using a regional dataset. It is well-known that the missing data outside the data area (the far-zone) has limited impact on the solution over the area of interest if the data are reduced for the contribution of a global gravity model. Hence, it is not necessary to modify the Poisson kernel in the case of the remove-compute-restore (RCR) scheme. In the case of RCR, a one-degree cap has been shown to be sufficient (Huang 2002).

After selecting an inner zone cap size, it remains to discretize the integral and solve the linear system of equations. The inner zone (\(\psi <{\psi }_{0}\)) contribution of the (residual) gravity disturbance at point \(p\) can be expressed by numerical discretization of the integral in Eq. (40) as.

$$ {\updelta }g_{i}^{p} = \mathop \sum \limits_{j = 1}^{M} B_{ij} {\updelta }g_{j}^{q^{\prime}} $$
(41)

where

$$ \begin{aligned} & B_{ij} \\ & = \left\{ {\begin{array}{ll} {\frac{R}{{4\pi r_{i} }}\sigma_{j} \wp \left( {R,{{\cos}}(\psi_{ij} ),r_{p} } \right)} & {{\text{if}} \psi_{ij} \le \psi_{0} ;{\text{and}} i \ne j} \\ {\frac{R}{r}d^{l} \left( {r_{p} ,\psi_{0} ,R} \right) - \mathop \sum \limits_{j = 1,j \ne i}^{K} B_{ij} } & {{\text{if}} \psi_{ij} \le \psi_{0} ;{\text{and }}i = j} \\ 0 & {{\text{if}}\; \psi_{ij} > \psi_{0} } \\ \end{array} } \right.\end{aligned} $$
(42)
$$ d^{l} \left( {r_{p} ,\psi_{0} ,R} \right) = \frac{1}{2}\left[ {\frac{r + R}{r}\left( {1 - \frac{r - R}{l}} \right)} \right] $$
(43)

and the integral area on a unit sphere is \({\sigma }_{j}={\Delta }_{\phi }{\Delta }_{\lambda }\mathrm{cos}(\phi )\).

Eq. (40) can be expressed in a matrix representation as \(={\varvec{g}}\), which is then solved by direct computations, or the following iterations (Huang 2002):

$$ {\varvec{f}} = A{\varvec{f}} + {\varvec{g}} $$
(44)

where \(A=I-B\), I being an identity matrix. Eq. (44) shows the sums of the observed downward continued gravity disturbances and correction terms. The iteration in Eq. (44) can start from

$$ {\varvec{f}}_{{}}^{0} = {\varvec{g}} $$
(45)

Followed by, \({{\varvec{f}}}^{1}={\varvec{A}}{{\varvec{f}}}^{0}+{\varvec{g}}\), \({{\varvec{f}}}^{2}={\varvec{A}}{{\varvec{f}}}^{1}+{\varvec{g}}\),…,\({{\varvec{f}}}^{{\varvec{i}}}={\varvec{A}}{{\varvec{f}}}^{{\varvec{i}}-1}+{\varvec{g}}\), and \({{\varvec{f}}}^{{\varvec{i}}+1}={\varvec{A}}{{\varvec{f}}}^{{\varvec{i}}}+{\varvec{g}}\).

The dimension of the linear system in Eq. (44) depends on the discretization step and data coverage, and can become very large. In order to avoid solving a huge system of linear equations, a block-wise approach is adopted to solve the DC for a block of 1° × 1° at a time without overlapping. No tile pattern is found in the DC results. Each side of the block is extended at least 1° in spherical distance to avoid edge effects. The iteration stops when the RMS differences between two consecutive solutions are smaller than a specified threshold value. The RMS differences correspond to the solution residuals of the previous iteration, as shown by Kingdon and Vaníček (2011). In this study, the threshold RMS difference and residual are set to 0.001 and 1.0 mGal, respectively, for the convergence of the inverse Poisson method.

1.1.6 Moritz’s analytical downward continuation

Moritz’s ADC method (Moritz 1980, Sect. 45) was originally derived to continue the gravity anomaly from the telluroid to the point level for the determination of the height anomaly by Molodensky’s theory. The DC correction is expressed as the sum of a series of terms. The G1 term represents the first order approximation, i.e., a linear gradient DC correction. When the DC of the data is well-posed, the higher-order terms attenuate rapidly to lead to a convergent correction. Otherwise the series becomes numerically divergent when the DC problem is numerically ill-posed, and must be truncated empirically to produce an approximate result.

In this study, we apply it to the DC of airborne gravity disturbance. The computational equations (Huang 2002, Sect. 5.2) are shown below.

$$ \delta g\left( {r_{g} ,\phi_{p} ,\lambda_{p} } \right) = \mathop \sum \limits_{n = 0}^{\infty } G_{n} $$
(46)
$$ G_{0} \left( {r_{p} ,\phi_{p} ,\lambda_{p} } \right) = \delta g\left( {r_{p} ,\phi_{p} ,\lambda_{p} } \right) $$
(47)
$$ G_{n} \left( {r_{p} ,\phi_{p} ,\lambda_{p} } \right) = - \mathop \sum \limits_{r = 1}^{n} H^{r} L_{r} \left( {G_{n - r} \left( {r_{p} ,\phi_{p} ,\lambda_{p} } \right)} \right) $$
(48)

where the \(L\) operator is given by:

$$\begin{aligned}& L\left( {G\left( {r_{p} ,\phi_{p} ,\lambda_{p} } \right)} \right) = \frac{{R^{2} }}{2\pi }\iint\limits_{{\sigma_{R}^{{}} }} {\frac{G\left( Q \right) - G\left( P \right)}{{(2R\sin \left( {\frac{\psi }{2}} \right))^{3} }}d\sigma_{R} \left( Q \right)} \\ & \quad {- \frac{1}{R}G\left( {r_{p} ,\phi_{p} ,\lambda_{p} } \right)} \end{aligned}$$
(49)

Empirically the DC is truncated to order 5 so that a potential divergence manifest itself, and can be controlled numerically for an ill-posed case.

Appendix B

2.1 Method for computing a geoid model from band-limited downward continued airborne gravity data

Using the remove-compute-restore (RCR) Stokes–Helmert scheme, a geoid model is computed from the gravity disturbances given on the surface of the reference ellipsoid by (Huang and Véronneau 2013):

$$ N\left( {\Omega } \right) = N_{0} \left( {\Omega } \right) + N_{1} \left( {\Omega } \right) + N_{{\text{H}}} \left( {\Omega } \right) + \delta N_{{{\text{PITE}}}} \left( {\Omega } \right) $$
(50)

where \({N}_{0}\left(\Omega \right) \mathrm{and} {N}_{1}\left(\Omega \right)\) are degree-0 and -1 terms; the last term stands for the primary indirect topographical effect on the geoid, and the Helmert co-geoid components above degree-1 can be expressed as

$$ N_{{\text{H}}} \left( {\Omega } \right) = N_{{{\text{HRef}}}} \left( {\Omega } \right) + dN\left( {\Omega } \right) $$
(51)
$$ N_{{{\text{HRef}}}} \left( {\Omega } \right) = \zeta_{{{\text{Ref}}}} \left( {\Omega } \right) - \delta N_{{{\text{DTE}}}} \left( {\Omega } \right) $$
(52)
$$ dN\left( {\Omega } \right) = \frac{R}{{4\pi \gamma \left( {\Omega } \right)}}\int\limits_{{{\upsigma }_{{0}} }} {S_{{{\text{MDB}}}} \left( \psi \right)dg_{{\text{H}}} \left( {{\Omega }^{^{\prime}} } \right)d\sigma } $$
(53)

When the residual geoid height is determined to the same spectral band as the reference field, the direct topographical effect (DTE) on the airborne data are computed to the same band as the reference field. Then we have

$$\begin{aligned} dg_{{\text{H}}} \left( {\Omega } \right) &= \left[ {\Delta g_{{{\text{Airborne}}}} \left( {\Omega } \right) - \delta \Delta g_{{{\text{DTE}}}} \left( {\Omega } \right)} \right] \\&- \left[ {\Delta g_{{{\text{Ref}}}} \left( {\Omega } \right) - \Delta g_{{{\text{DTE}}}} \left( {\Omega } \right)} \right] \end{aligned}$$
(54)

or

$$ dg_{{\text{H}}} \left( {\Omega } \right) = \Delta g_{{{\text{Airborne}}}} \left( {\Omega } \right) - \Delta g_{{{\text{Ref}}}} \left( {\Omega } \right) $$
(55)

Therefore, Eq.(53) can be re-written as

$$ \begin{aligned} dN\left( {\Omega } \right) &= d\zeta \left( {\Omega } \right) = \frac{R}{{4\pi \gamma \left( {\Omega } \right)}}\int\limits_{{{\upsigma }_{0} }} {S_{{{\text{MDB}}}} \left( \psi \right)}\\ & \quad{ \times \left[ {\Delta g_{{{\text{Airborne}}}} \left( {\Omega } \right) - \Delta g_{{{\text{Ref}}}} \left( {\Omega } \right)} \right]d\sigma } \end{aligned}$$
(56)

Considering Eq. (5156) and the degree-0 and degree-1 terms for the geoid height are equal to the same terms for the height anomaly on these reference ellipsoid, Eq. (50) can be re-written as

$$ N\left( {\Omega } \right) = \zeta_{0} \left( {\Omega } \right) + \zeta_{1} \left( {\Omega } \right) + \zeta_{{{\text{Ref}}}} \left( {\Omega } \right) + d\zeta \left( {\Omega } \right) + C_{{\text{T}}} \left( {\Omega } \right) $$
(57)

where

$$ C_{{\text{T}}} \left( {\Omega } \right) = - \delta N_{{{\text{DTE}}}} \left( {\Omega } \right) + \delta N_{{{\text{PITE}}}} \left( {\Omega } \right) $$
(58)

The \({C}_{\text{T}}\left(\Omega \right)\) term is given in (Huang and Véronneau 2015, equations A910), which converts the height anomaly evaluated on the reference ellipsoid to the geoid height. Note that Sjöberg (Eq. 23, 2007) derived an equivalent equation to (10) to the 3rd order approximation, and named it the topographic bias with a negative sign.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Li, X., Huang, J., Klees, R. et al. Characterization and stabilization of the downward continuation problem for airborne gravity data. J Geod 96, 18 (2022). https://doi.org/10.1007/s00190-022-01607-y

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1007/s00190-022-01607-y

Keywords

Navigation