Skip to main content
Log in

Forward calculation of gravity and its gradient using polyhedral representation of density interfaces: an application of spherical or ellipsoidal topographic gravity effect

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

Abstract

A density interface modeling method using polyhedral representation is proposed to construct 3-D models of spherical or ellipsoidal interfaces such as the terrain surface of the Earth and applied to forward calculating gravity effect of topography and bathymetry for regional or global applications. The method utilizes triangular facets to fit undulation of the target interface. The model maintains almost equal accuracy and resolution at different locations of the globe. Meanwhile, the exterior gravitational field of the model, including its gravity and gravity gradients, is obtained simultaneously using analytic solutions. Additionally, considering the effect of distant relief, an adaptive computation process is introduced to reduce the computational burden. Then features and errors of the method are analyzed. Subsequently, the method is applied to an area for the ellipsoidal Bouguer shell correction as an example and the result is compared to existing methods, which shows our method provides high accuracy and great computational efficiency. Suggestions for further developments and conclusions are drawn at last.

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

Similar content being viewed by others

References

  • Amante C, Eakins BW (2009) Etopo1 1arc-minute global relief model: procedures, data sources and analysis. In: NOAA Technical Memorandum NESDIS NGDC-24 Boulder, CO

  • An YL, Guo LH, Zhang MH (2015) High precision computation and numerical value characteristics of gravity emendation values arising from mass of the earth’s crust at the distance over 169 km from the observation point. Geophys Geochem Explor 39(1):1–11 (in Chinese)

    Google Scholar 

  • Ardalan AA, Safari A (2004) Ellipsoidal terrain correction based on multi-cylindrical equal-area map projection of the reference ellipsoid. J Geod 78:114–123

    Google Scholar 

  • Asgharzadeh MF, Frese RRB, Kim HR, Leftwich TE, Kim JW (2007) Spherical prism gravity effects by Gauss–Legendre quadrature integration. Geophys J Int 169:1–1

    Article  Google Scholar 

  • Balmino G, Vales N, Bonvalot S, Briais A (2011) Spherical harmonic modelling to ultra-high degree of Bouguer and isostatic anomalies. J Geod 86:499–520

    Article  Google Scholar 

  • Bouman J, Ebbing J, Fuchs M (2013) Reference frame transformation of satellite gravity gradients and topographic mass reduction. J Geophys Res 118(2):759–774

    Article  Google Scholar 

  • Bullard EC (1936) Gravity measurements in East Africa. Philos Trans R Soc 235:445–534

    Article  Google Scholar 

  • Çavsak H (2012) The effects of the Earth’s curvature on gravity and geoid calculations. Pure Appl Geophys 169(4):733–740

    Article  Google Scholar 

  • Chapin DA (1996) The theory of the Bouguer gravity anomaly: a tutorial. Lead Edge 15:361–363

    Article  Google Scholar 

  • Claessens SJ, Hirt C (2013) Ellipsoidal topographic potential: new solutions for spectral forward gravity modeling of topography with respect to a reference ellipsoid. J Geophys Res 118(11):5991–6002

    Article  Google Scholar 

  • Du JS, Chen C, Liang Q, Wang LS, Zhang Y, Wang QG (2012) Gravity anomaly calculation based on volume integral in spherical cap and comparison with the Tesseroid–Taylor series expansion approach. Acta Geodaetica Cartogr Sin 41(3):339–346 (in Chinese)

    Google Scholar 

  • Dutton GH (1999) A hierarchical coordinate system for geoprocessing and cartography. In: Lecture Notes in Earth Sciences, vol. 79. Springer, Berlin

  • Eshagh M (2013) Numerical aspects of EGM08-based geoid computations in Fennoscandia regarding the applied reference surface and error propagation. J Appl Geophys 96:28–32

    Article  Google Scholar 

  • Eshagh M, Sjöberg LE (2009) Atmospheric effect on satellite gravity gradiometry data. J Geodyn 47:9–19

    Article  Google Scholar 

  • Fairhead JD, Green CM, Blitzkow D (2003) The use of GPS in gravity surveys. Lead Edge 22(10):954–959

    Article  Google Scholar 

  • Feather WE, Dentith MC (1997) A geodetic approach to gravity data reduction for geophysics. Comput Geosci 23(10):1063–1070

  • Fekete G, Treinish L (1990) Sphere quadtrees: A new data structure to support the visualization of spherically distributed data. In: Extracting meaning from complex data: processing, display, interaction, SPIE, vol 1259, pp 242–250

  • Förste C, Bruinsma S, Abrikosov O, Flechtner F, Marty JC, Lemoine JM, Biancale R (2014) EIGEN-6C4—the latest combined global gravity field model including GOCE data up to degree and order 1949 of GFZ Potsdam and GRGS Toulouse. In: EGU general assembly conference abstracts, vol 16, May 2014

  • Fuller RB (1975) Synergetics. MacMillan, New York

    Google Scholar 

  • Gregory M (1999) Comparing inter-cell distance and cell wall midpoint criteria for discrete global grid systems. PhD thesis, Department of Geosciences, Oregon State University

  • Grombein T, Seitz K, Heck B (2013) Optimized formulas for the gravitational field of a tesseroid. J Geod 87(7):645–660

    Article  Google Scholar 

  • Heck B, Seitz K (2007) A comparison of tesseroid, prism and point-mass approaches for mass reductions in gravity field modeling. J Geophys 81:121–136

    Google Scholar 

  • Hensel EG (1992) Discussion on: ‘An exact solution for the gravity curvature (Bullard B) correction’ by T. R. Lafehr. Geophysics 57:1093–1094

    Article  Google Scholar 

  • Hirt C, Kuhn M, Featherstone WE, Gottl F (2012) Topographic/isostic evaluation of new-generation GOCE gravity field models. J Geophys Res 117(B5):B05407. doi:10.1029/2011JB008878

    Article  Google Scholar 

  • Kenner H (1976) Geodesic math and how to use it. University of California Press, Berkeley

    Google Scholar 

  • Lafehr TR (1991a) An exact solution for the gravity curvature (Bullard B) correction. Geophysics 56(8):1179–1184

    Article  Google Scholar 

  • Lafehr TR (1991b) Standardisation in gravity reduction. Geophysics 56:1170–1178

    Article  Google Scholar 

  • Lafehr TR (1992) Discussion on: ‘An exact solution for the gravity curvature (Bullard B) correction’ by T. R. Lafehr. Geophysics 57:1094

    Article  Google Scholar 

  • Lafehr TR (1998) On Talwani’s ‘Errors in the total Bouguer reduction’. Geophysics 63:1131–1136

    Article  Google Scholar 

  • Lane R (2009) Some issues and insights for gravity and magnetic modeling at the region to continent scale. ASEG Ext Abstr 2009(1):1–11

  • Leaman DE (1998) The gravity terrain correction—practical considerations. Explor Geophys 29:467–471

    Article  Google Scholar 

  • Lemoine, FG, Kenyon SC, Factor JK, Trimmer RG, Pavlis NK, Chinn DS, Cox CM, Klosko SM, Luthcke SB, Torrence MH, Wang YM, Williamson RG, Pavlis EC, Rapp RH, Olson TR (1998) The Development of the Joint NASA GSFC and the National Imagery and Mapping Agency (NIMA) Geopotential Model EGM96. NASA/TP-1998-206861, July 1998

  • Li CY, Sideris MG (1994) Improved gravimetric terrain corrections. Geophys J Int 119:740–752

    Article  Google Scholar 

  • Li X, Götze H (2001) Ellipsoid, geoid, gravity, geodesy, and geophysics. Geophysics 66(6):1660–1668

    Article  Google Scholar 

  • Mikuška J, Pašteka R, Marušiak I (2006) Estimation of distant relief effect in gravimetry. Geophysics 71(6):J59–J69

    Article  Google Scholar 

  • Nowell DAG (1999) Gravity terrain corrections—an overview. J Appl Geophys 42:117–134

    Article  Google Scholar 

  • Parker RL (1973) The rapid calculation of potential anomalies. Geophys J R Astron Soc 31(4):447–455

    Article  Google Scholar 

  • Parker RL (1975) Improved fourier terrain correction, Part 1. Geophysics 60(4):1007–1017

    Article  Google Scholar 

  • Roussel C, Verdun J, Cali J, Masson F (2015) Complete gravity field of an ellipsoidal prism by Gauss–Legendre quadrature. Geophys J Int 203(3):2220–2236

    Article  Google Scholar 

  • Sadourny R, Arakawa A, Mintz Y (1968) Integration of the nondivergent barotropic vorticity equation with an icosahedral-hexagonal grid for the sphere. Mon Weather Rev 96(6):351–356

    Article  Google Scholar 

  • Sahr K, White D, Kimerling JA (2003) Geodesic discrete global grid systems. Cartogr Geogr Inf Sci 30(2):121–134

    Article  Google Scholar 

  • Sjöberg LE (2001) Topographic and atmospheric corrections of gravimetric geoid determination with special emphasis on the effects of harmonics of degrees zero and one. J Geod 75(5–6):283–290

    Google Scholar 

  • Talwani M (1998) Errors in the total Bouguer reduction. Geophysics 63:1125–1130

    Article  Google Scholar 

  • Tenzer R, Hamayun K, Vajda P (2008) Global secondary indirect effects of topography, bathymetry, ice and sediments. Contrib Geophys Geod 38(2):209–216

    Google Scholar 

  • Tenzer R, Hamayun K, Vajda P (2009) Global maps of the CRUST 2.0 crustal components stripped gravity disturbances. J Geophys Res 114:B05408

    Article  Google Scholar 

  • Thuburn J (1997) A PV-based shallow-water model on a hexagonal-icosahedral grid. Mon Weather Rev 125:2328–2347

    Article  Google Scholar 

  • Tsoulis D (1998) A combination method for computing terrain corrections. Phys Chem Earth 23(1):53–58

    Article  Google Scholar 

  • Tsoulis D (2012) Analytical computation of the full gravity tensor of a homogeneous arbitrarily shaped polyhedral source using line integrals. Geophysics 77(2):F1–F14

    Article  Google Scholar 

  • Tsoulis D, Novák P, Kadlec M (2009a) Evaluation of precise terrain effects using high-resolution digital elevation models. J Appl Geophys 114(B02404):1–14

    Google Scholar 

  • Tsoulis D, Jamet O, Verdum J, Gonindard N (2009b) Recursive algorithms for the computation of the potential harmonic coefficients of a constant density polyhedron. J Geod 83:925–942

    Article  Google Scholar 

  • Tsoulis D, Petrović S (2001) Short note: on the singularities of the gravity field of a homogeneous polyhedral body. Geophysics 66(2):535–539

    Article  Google Scholar 

  • Tsoulis D, Wziontek H, Petrović S (2003) A bilinear approximation of the surface relief in terrain correction computations. J Geod 77:338–344

    Article  Google Scholar 

  • Uieda L, Bomfim EP, Braiterberg C, Molina E (2011) Optimal forward calculation method of the Marussi tensor due to a geologic structure at GOCE height. In: Proceedings of the 4th international GOCE user workshop, ESA SP-696, July 2011, Munich

  • Uieda L, Barbosa VCF, Braitenberg C (2015) Tesseroids: forward-modeling gravitational fields in spherical coordinates. Geophysics 81(5):F41–F48

    Article  Google Scholar 

  • Urso MG (2013) On the evaluation of the gravity effects of polyhedral bodies and a consistent treatment of related singularities. J Geod 87:239–252

    Article  Google Scholar 

  • Wang HS, Wu P, Wang ZY (2006) Approach for spherical harmonic analysis of non-smooth data. Comput Geosci 32:1654–1668

    Article  Google Scholar 

  • Werner AR, Scheeres JD (1996) Exterior gravitation of a polyhedron derived and compared with harmonic and mascon gravitation representations of Asteroid 4769 Castalia. Celest Mech Dyn Astron 65(3):313–344

    Google Scholar 

  • White D, Kimerling JA, Overton SW (1992) Cartographic and geometric components of a global sampling design for environmental monitoring. Cartogr Geogr Inf Syst 19(1):5–22

    Google Scholar 

  • White D, Kimerling JA, Sahr K, Song L (1998) Comparing area and shape distortion on polyhedral-based recursive partitions of the sphere. Int J Geogr Inf Sci 12:805–827

    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(E1):1715–1724

    Article  Google Scholar 

  • Wild-Pfeiffer F (2008) A comparison of different mass elements for use in gravity gradiometry. J Geod 82:637–653

    Article  Google Scholar 

  • Williamson LD (1968) Integration of the barotropic vorticity equation on a spherical geodesic grid. Tellus 20(4):642–653

    Article  Google Scholar 

Download references

Acknowledgements

We thank Dr. Walter Mooney, Dr. Mikhail Kaban and three anonymous reviewers for their precious suggestions for preparing and revising the manuscript. This study is supported by the China Scholarship Council (No. 1609130003) and the Natural Science Foundation of China (No. 41574070). The source code of the method is available upon request.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Chao Chen.

Appendices

A Appendix

In order to map the partition on the icosahedral facets outward to a certain surface, geocentric radiuses of the surface under an equal-angle grid are firstly determined (Fig. 14). As for the terrain surface of the Earth, discussed by Balmino et al. (2011) and Tenzer et al. (2009), given some systematic approximations adopted, the geocentric radius R could be expressed as \(R=r+H+N\) without introducing large error, in which r means the geocentric radius of the Earth’s reference sphere or ellipsoid, H and N are the corresponding altitude and geoid undulation. If the radius r is latitude-dependent, the Earth’s ellipticity can be easily taken into account by \(r=ab\sqrt{{a}^{2}\hbox {sin}^{2}{\theta }_{c}{+b}^{2}\cos ^{2}{\theta }_{c}}\), where \(\theta _c \) is the geocentric colatitude, a and b are the semimajor and semiminor axes of the Earth’s reference ellipsoid, respectively.

Fig. 14
figure 14

A diagrammatic sketch of the relationship between the PGIM’s vertices and an equal-angle grid. In which, m is a vertex of the PGIM, and \(e_{1}\) to \(e_{4}\) are four data points of the equal-angle grid. The geocentric radius of m is interpolated by these four points’ radiuses

After determining geocentric radiuses under the equal angular grid, radiuses of the PGIM’s vertices are interpolated using the spherical bilinear interpolation method (Wang et al. 2006) as following. Geocentric radius \({r}_{ave}\) of a PGIM’s vertex \({m}\left( {\phi _{m}, {\theta }_{m} } \right) \) could be determined by grid points closed to m. Provided \(e_{1}(\phi _{1},\theta _{1},r_{1}),e_{2}(\phi _{2},\theta _{2},r_{2}),e_{3} (\phi _{3},\theta _{3},r_{3})\) and \(e_{4}(\phi {4},\theta {4},r_{4})\) are four data points of the equal angular grid which form the smallest spherical quadrangular region with m in it (Fig. 14). The interpolation can be performed as:

$$\begin{aligned}&r_{\mathrm {ave}} \left( {\phi _{m} ,\theta _{m} } \right) =\left( \hbox {Acos}\phi _{m} +B\theta _{m} +C\theta _{m} \cos \phi _{m} +D \right) /\varDelta \nonumber \\ \end{aligned}$$
(1)
$$\begin{aligned}&{A} =\Delta {\theta }_{1} \left( {r}_{4} {-r}_{2} \right) +{\theta }_{2} \left( {r}_{1} {-r}_{3} \right) \end{aligned}$$
(2)
$$\begin{aligned}&{B}=\left( {r}_{1} {-r}_{2} \right) \cos \phi _{1} +\left( {r}_{4} {-r}_{3} \right) \cos \phi _{2} , \end{aligned}$$
(3)
$$\begin{aligned}&{C}= r_{3}+r_{2} {-r}_{1} {-r}_{4} , \end{aligned}$$
(4)
$$\begin{aligned}&{D}=\theta _{2} \cos \phi _{2} {r}_{3} {-\theta }_{2} \cos \phi _{1} {r}_{1}+\theta _{1} \cos \phi _{1} {r}_{2}\nonumber \\&\qquad \quad -\,\theta _{1} \cos \phi _{2} {r}_{4} , \end{aligned}$$
(5)
$$\begin{aligned}&\varDelta =\left( {\theta }_{2} {-\theta }_{1} \right) \left( {\cos \phi _{2} -\cos \phi _{1} } \right) . \end{aligned}$$
(6)

Discussed by Du et al. (2012), the interpolation could be taken as an average and error reduction process of the equal angular data set. The error of the interpolation is very low if the resolution of the original grid data is very high. And Tsoulis et al. (2003) pointed out that the resulting surfaces had the geometrical constraint to pass through the exact positions of the original data. Therefore, the error introduced by the interpolation is ignored here.

B Appendix

The exterior gravitational potential, along with its vectors and gradient tensors, of a polyhedron can be forward calculated using the algorithm proposed by Werner and Scheeres (1996). The algorithm is derived in closed form so that there is no error produced during the process of forward calculating theoretically. Assuming the distribution of density inside a polyhedron, expressed as \(\rho \) in the following equations, is constant, all facets of the polyhedron are triangles and vertices of a facet are ordered anticlockwise as one observes the facet from outside of the polyhedron. Then the polyhedron’s exterior gravitational potential V, gradient \(\nabla {V}\) and gradient tensors \(\nabla \nabla {V}\) are given as below, in which G is the gravitational constant. Gradient directions here are x, y and z in the Cartesian coordinate system.

$$\begin{aligned}&V=\frac{1}{2}{G}\rho \left( {\sum \nolimits _{e=1}^{n}} \vec {r}_{e}\cdot {\mathbf{E}}_{e} \cdot \vec {r}_{e} \cdot {L}_{e}\right. \nonumber \\&\qquad \left. +{\sum \nolimits _{f=1}^{m}} \vec {r}_{f}\cdot {\mathbf{F}}_{f}\cdot \vec {{r}_{f} }\cdot {\omega }_{f} \right) , \end{aligned}$$
(7)
$$\begin{aligned}&\nabla V = -G \rho \left( {\sum \nolimits _{e=1}^{n}{\mathbf{E}}_{e}\cdot \vec {{r}_{e} }\cdot L_{e} +\sum \nolimits _{{f=1}}^{m} {\mathbf{F}}_{f} \cdot \vec {r}_{f} \cdot {\omega }_{f} } \right) , \nonumber \\ \end{aligned}$$
(8)
$$\begin{aligned}&\nabla \nabla V = -G\rho \left( \sum \nolimits _{e=1}^{n} {\mathbf{E}}_{e} \cdot {L}_{e} +\sum \nolimits _{f=1}^{m} {\mathbf{F}}_{f} \cdot {\omega }_{f} \right) . \end{aligned}$$
(9)

in which m and n represent number of facets and edges of the polyhedron, respectively. This feature indicates that the exterior gravitational fields of a polyhedron could be expressed as a sum of effect of all its facets and edges. For instance, as the sketch of a polyhedron shown in Fig. 15, \(\vec {{r}_{f} }\) and \(\vec {{r}_{e} }\) stand for two vectors from the observation point P to a triangular facet of the polyhedron, ABC, and one of its edges, CA. Meanwhile, \({\mathbf{F}}_{f} \) and \({\mathbf{E}}_{e} \) represent tensor products of facet and edge accordingly. \({\omega }_{f} \) and \({L}_{e} \) are two coefficients decided by the relative position between the observing point P and the face and edge. The expressions mentioned above are given by

Fig. 15
figure 15

A diagrammatic sketch of a tetrahedron ABCD placed in the Cartesian coordinate. P is the observation point, \(\vec {r}\) represents vector, and \(\vec {n}\) is a normal vector

$$\begin{aligned} {\mathbf{E}}_{e}= & {} {\hat{n}}_{1} {\hat{n}}_{ca} +{\hat{n}}_{2} {\hat{n}}_{ac},\end{aligned}$$
(10)
$$\begin{aligned} {\mathbf{F}}_{f}= & {} {\hat{n}}_{f} {\hat{n}}_{f} , \end{aligned}$$
(11)
$$\begin{aligned} {L}_{e}= & {} \frac{{\bar{r}}_{i} +{\bar{r}}_{j}+ d_{ij}}{{r}_{i} + {r}_{j} {-d}_{ij}} , \end{aligned}$$
(12)
$$\begin{aligned} \omega _{f}= & {} 2\tan ^{-1}\nonumber \\&\times \frac{\vec {r}_{i} \cdot \vec {r}_{j}\times \vec {r}_{k}}{{\bar{r}_{i}}{\bar{r}}_{j} \bar{r}_{k}+ \bar{r}_{i}\left( {\vec {r}_{j} \cdot \vec {r}_{k} } \right) + \bar{r}_{j} \left( {\vec {r}_{k} \cdot \vec {r}_{i} } \right) + \bar{r}_{k} \left( {\vec {r}_{i} \cdot \vec {r}_{j} } \right) }.\nonumber \\ \end{aligned}$$
(13)

In equations above, \(\vec {r}_{i}\), \(\vec {r}_\mathrm{j} \) and \(\vec {r}_\mathrm{k} \) represent three vectors from P to the face’s vertices C, A and B, with its module \(\bar{r}_{i}\), \(\bar{r}_{j}\) and \(\bar{r}_{k}\), respectively. \(\vec {n}_\mathrm{f} \) stands for the outside normal vector of the face ABC. It equals the normalized vector of \(\overrightarrow{AB}\times \overrightarrow{BC}\). \(d_{ij} \) represents the distance between the ends of vectors \(\vec {r}_\mathrm{i}\) and \(\vec {r}_{j}\) which equals the length of edge CA here. Moreover, \(\overrightarrow{n}_{1}\) and \(\overrightarrow{n}_{2}\) are two outside normal vectors of edge CA that lie on the face ABC and face ACD, respectively. These two vectors could be obtained by the cross-product between a face’s outside normal vector and one of its edge’s normal vector accordingly. For instance, \(\overrightarrow{n}_{1}\) equals \(\overrightarrow{n}_{f} \times \overrightarrow{n}_{ca}\). Expressions such as \(\mathbf{F}_{f} ={\hat{n}} _{f} {\hat{n}} _\mathrm{f}\) means the tensor product of two normal vectors which is a \(3\times 3\) matrix. For each facet of the polyhedron, it has a \({\mathbf{F}}_{f}\) and \(\omega _{f} \) associated with. And each edge has an individual \({\mathbf{E}}_e \) and \({\mathbf{L}}_{e} \). Subsequently, the gravitational field of the polyhedron could be obtained by calculating the sum of effects of all its facets and edges.

As all computations are conducted in the Cartesian coordinate, the outcome gravitational fields are with respect to the Cartesian coordinate as well. A coordinate transformation matrix R is added into Eqs. (8) and (9), as we need to forward calculate the gravitational fields of a polyhedron in directions reference to the spherical coordinate. Given the position of the observation point P in spherical coordinates \(\left( {\phi _P ,\theta _P ,\bar{r}_{P}}\right) \), then R is given by

$$\begin{aligned} {\mathbf{R}}=\left| {{\begin{array}{ccc} {\sin \theta _P \cos \phi _P }&{}\quad {\sin \theta _P \sin \phi _P }&{}\quad {\cos \theta _P } \\ {\cos \theta _P \cos \phi _P }&{}\quad {\cos \theta _P \sin \phi _P }&{}\quad {-\hbox {sin}\theta _P } \\ {-\hbox {sin}\phi _P }&{}\quad {\cos \phi _P }&{}\quad {0} \\ \end{array}}} \right| . \end{aligned}$$
(14)

Subsequently, gradient and gradient tensors of a polyhedron’s gravitational potential oriented in the directions \(\phi \), \(\theta \) and r are expressed as below.

$$\begin{aligned} \nabla V&=-G\rho \left( \sum \nolimits _{e=1}^n {\mathbf{E}}_{e} \cdot \vec {r}_{e}\cdot {\mathbf{R}}^\mathrm{T}\cdot {L}_{e} \right. \nonumber \\&\qquad \left. +\,\sum \nolimits _{f=1}^{m} {\mathbf{F}}_{f}\cdot \vec {r}_{f} \cdot {\mathbf{R}}^\mathrm{{T}}\cdot \omega _{f} \right) , \end{aligned}$$
(15)
$$\begin{aligned} \nabla \nabla V&=-G\rho \left( \sum \nolimits _{{e=1}}^{n} {\mathbf{R}}\cdot {\mathbf{E}}_{e}\cdot {\mathbf{R}}^\mathrm{{T}}\cdot {L}_{e} \right. \nonumber \\&\qquad \left. +\,\sum \nolimits _{f=1}^{m} {\mathbf{R}}\cdot {\mathbf{F}}_{f} \cdot {\mathbf{R}}^\mathrm{{T}}\cdot \omega _{f} \right) . \end{aligned}$$
(16)

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Zhang, Y., Chen, C. Forward calculation of gravity and its gradient using polyhedral representation of density interfaces: an application of spherical or ellipsoidal topographic gravity effect. J Geod 92, 205–218 (2018). https://doi.org/10.1007/s00190-017-1057-3

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00190-017-1057-3

Keywords

Navigation