Skip to main content
Log in

The effect of anomalous global lateral topographic density on the geoid-to-quasigeoid separation

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

Abstract

The geoid can be computed from the quasigeoid by applying the geoid-to-quasigeoid separation. The geoid-to-quasigeoid separation is also needed for a vertical datum unification. Information about the actual topographic density distribution is required to determine accurately the geoid and orthometric heights. In this study, we estimate the effect of (lateral) anomalous topographic density on the geoid-to-quasigeoid separation. This became possible after releasing the first global lateral topographic density model UNB_TopoDens. This model provides also information about topographic density uncertainties. According to our estimates by using the UNB_TopoDens model, the effect of anomalous topographic density on the geoid-to-quasigeoid separation globally varies mostly within ±0.02 m. In mountainous regions (particularly in the Himalayas and Tibet), this effect could reach (or even exceed) ±0.1 m. The analysis also reveals that the errors in computed values of the geoid-to-quasigeoid separation attributed to the UNB_TopoDens density uncertainties (provided in terms of standard deviations for representative lithologies) are globally mostly within a few centimeters. In parts of mountainous regions with large topographic density uncertainties, however, these errors might exceed ±0.1 m. There is another crucial aspect we address here. According to the UNB_TopoDens model, the average topographic density for the whole landmass (except for polar glaciers) is 2247 kg m−3. This average density is considerably smaller than the value of 2670 kg m−3 that is typically used in geodetic and geophysical applications. This new estimate, if confirmed independently, might have implications on Helmert orthometric heights (adopted in many countries for a vertical datum realization). Changes in Helmert orthometric heights due to adopting this new estimate of the average topographic density are systematic and reach several decimeters. We, therefore, propose to use a gravimetric geoid model for a practical realization of vertical datums that incorporates the topographic density information in countries where Helmert orthometric heights are adopted. This recommendation is fully compatible with modern concepts for a vertical datum realization based on using a gravimetric geoid model and geodetic (ellipsoidal) heights. We also address inconsistencies in computations of Helmert orthometric heights and gravimetric geoid models, and propose to use only accurately computed orthometric heights (including variable topographic density term) to combine (or fit) a gravimetric geoid model with geometric geoid heights at GPS-leveling benchmarks.

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

Similar content being viewed by others

Data availability

The global geopotential model EIGEN-6C4 is freely available via ICGEM (International Centre for Global Earth Models) website. The digital elevation model Earth2014 SUR is publicly accessible via the following link http://ddfe.curtin.edu.au/models/Earth2014/.

The digital topographic density model UNB_TopoDens is publicly accessible via the following link: https://www.unb.ca/fredericton/engineering/depts/gge/resources.html

References

  • Artemjev ME, Kaban MK, Kucherinenko VA, Demjanov GV, Taranov VA (1994) Subcrustal density inhomogeneities of northern Eurasia as derived from the gravity data and isostatic models of the lithosphere. Tectonophysics 240:248–280

    Article  Google Scholar 

  • Bear GW, Al-Shukri HJ, Rudman AJ (1995) Linear inversion of gravity data for 3-D density distributions. Geophysics 60(5):1354–1364

    Article  Google Scholar 

  • Cardenas C (2018) The Earth Geological Globe: a 3D geological map of the earth constructed from 2D data. GeoExPro 15(3)

  • Carmichael RS (1989) Practical handbook of physical properties of rocks and minerals. CRC Press, Boca Raton

    Google Scholar 

  • Cutnell JD, Kenneth WJ (1995) Physics, 3rd edn. Wiley, New York

    Google Scholar 

  • Dziewonski AM, Anderson DL (1981) Preliminary reference Earth model. Phys Earth Planet Inter 25(4):297–356

    Article  Google Scholar 

  • Flury J, Rummel R (2009) On the geoid-quasigeoid separation in mountain areas. J Geod 83:829–847

    Article  Google Scholar 

  • Foroughi I, Tenzer R (2017) Comparison of different methods for estimating the geoid-to-quasigeoid separation. Geophys J Int 210(2):1001–1020

    Article  Google Scholar 

  • Foroughi I, Vaníček P, Kingdon RW, Goli M, Sheng M, Afrasteh Y, Novák P, Santos MC (2019) Sub-centimetre geoid. J Geodesy 93(6):849–868

    Article  Google Scholar 

  • Förste C, Bruinsma SL, Abrikosov O, Lemoine J-M, Schaller T, Götze H-J, Ebbing J, Marty J-C, Flechtner F, Balmino G, Biancale R (2014) EIGEN-6C4: the latest combined global gravity field model including GOCE data up to degree and order 2190 of GFZ Potsdam and GRGS Toulouse; presented at the 5th GOCE User Workshop, Paris, 25-28 November 2014

  • Fotopoulos G (2005) Calibration of geoid error models via a combined adjustment of ellipsoidal, orthometric and gravimetric geoid height data. J Geod 79:111–123

    Article  Google Scholar 

  • Fotopoulos G, Kotsakis C, Sideris MG (2003) How accurately can we determine orthometric height differences from GPS and geoid data. J Surv Eng 1:1–10

    Article  Google Scholar 

  • Grebenitcharsky RS, Rangelova EV, Sideris MG (2005) Transformation between gravimetric and GPS/levelling-derived geoids using additional gravity information. J Geodyn 39:527–544

    Article  Google Scholar 

  • Hartmann J, Moosdorf N (2012) The new global lithological map database GLiM: a representation of rock properties at the Earth surface. Geochem Geophys Geosyst 13(431):Q12004

    Google Scholar 

  • Heiskanen WH, Moritz H (1967) Physical geodesy. WH Freeman and Co, San Francisco

    Google Scholar 

  • Helmert FR (1884) Die mathematischen und physikalischen Theorien der höheren Geodäsie, vol 2. Teubner, Leipzig

    Google Scholar 

  • Helmert FR (1890) Die Schwerkraft im Hochgebirge, insbesondere in den Tyroler Alpen. Veröff Königl Preuss Geod Inst, No 1

  • Hewitt E, Hewitt RE (1979) The Gibbs-Wilbraham phenomenon: an episode in Fourier analysis. Arch Hist Exact Sci 21(2):129–160

    Article  Google Scholar 

  • Hinze WJ (2003) Bouguer reduction density, why 2.67? Geophysics 68(5):1559–1560

    Article  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(9):729–744

    Article  Google Scholar 

  • Hirt C, Kuhn M (2012) Evaluation of high-degree series expansions of the topographic potential to higher-order powers. J Geophys Res 117:B12407

    Google Scholar 

  • Hirt C, Rexer M (2015), Earth2014: 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, accepted for publication

  • Klees R, Prutkin I (2010) The combination of GNNS-levelling data and gravimetric (quasi-) geoid heights in the presence of noise. J Geod 84:731–749

    Article  Google Scholar 

  • Kuhtreiber N (1998) Precise geoid determination using a density variation model. Phys Chem Earth 23:59–63

    Article  Google Scholar 

  • Laske G, Masters G, Ma Z, Pasyanos ME (2012) CRUST1.0: an updated global model of Earth’s CRUST. In: Geophysical Research Abstract. EGU2012–37431

  • Ledersteger K (1968) Astronomische und Physikalische Geodäsie (Erdmessung). In: Jordan W, Eggert E, Kneissl M (eds) Handbuch der Vermessungskunde, vol V. Metzler, Stuttgart

    Google Scholar 

  • Listing JB (1873) Über unsere jetzige Kenntniss der Gestalt und Grösse der Erde, Nachrichten von der Königl. Gesellschaft der Wissenschaften und der GA, Universität zu Göttingen 3:33–98

    Google Scholar 

  • Mader K (1954) Die orthometrische Schwerekorrektion des Präzisions-Nivellements in den Hohen Tauern. Österreichische Zeitschrift für Vermessungswesen, Sonderheft, p 15

    Google Scholar 

  • Mankhemthong N, Doser DI, Baker MR (2012) Practical Estimation of Near-surface Bulk Density Variations Across the Border Ranges Fault System, Central Kenai Peninsula, Alaska. J Environ Eng Geophys 17(3):51–158

    Article  Google Scholar 

  • Martinec Z (1993) Effect of lateral density variations of topographical masses in view of improving geoid model accuracy over Canada. Contract report for Geodetic Survey of Canada, Ottawa, Canada

  • Martinec Z, Vaníček P, Mainville A, Veronneau M (1995) The effect of lake water on geoidal height. Manuscr Geod 20:193–203

    Google Scholar 

  • Molodensky MS (1945) Fundamental Problems of Geodetic Gravimetry (in Russian). TRUDY Ts NIIGAIK, 42, Geodezizdat, Moscow

  • Molodensky MS (1948) External gravity field and the shape of the Earth surface. Izv CCCP, Moscow (in Russian)

    Google Scholar 

  • Molodensky MS, Yeremeev VF, Yurkina MI (1960) Methods for Study of the External Gravitational Field and Figure of the Earth. TRUDY Ts NIIGAiK, Vol. 131, Geodezizdat, Moscow. English translation: Israel Program for Scientific Translation, Jerusalem 1962

  • Moritz H (2000) Geodetic Reference System 1980. J Geod 74:128–162

    Article  Google Scholar 

  • Nettleton LL (1939) Determination of density for reduction of gravitimeter observations. Geophysics 4:8176–8183

    Article  Google Scholar 

  • Niethammer T (1932) Nivellement und Schwere als Mittel zur Berechnung wahrer Meereshöhen. Schweizerische Geodätische Kommission

  • Niethammer T (1939) Das astronomische Nivellement im Meridian des St Gotthard, Part II, Die berechneten Geoiderhebungen und der Verlauf des Geoidschnittes. Astronomisch-Geodätische Arbeiten in der Schweiz, Vol 20, Swiss Geodetic Commission

  • Parasnis DS (2012) A study of rock densities in English Midlands. Geophys J Int 1952(6):252–271

    Google Scholar 

  • Pizzetti P (1911) Sopra il calcolo teorico delle deviazioni del geoide dall` ellissoide. Atti R Accademia delle Scienze di Torino, Vol 46

  • Prutkin I, Klees R (2008) The non-uniqueness of local quasi-geoids computed from terrestrial gravity anomalies. J Geod 82(3):147–156

    Article  Google Scholar 

  • Rathnayake S, Tenzer R, Novák P, Pitoňák M (2020) Effect of the lateral topographic density distribution on interpretational properties of Bouguer gravity maps. Geophys J Int 220(2):892–909

    Google Scholar 

  • 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 37:1035–1074

    Article  Google Scholar 

  • Santos MC, Vaníček P, Featherstone WE, Kingdon R, Ellmann A, Martin B-A, Kuhn M, Tenzer R (2006) The relation between rigorous and Helmert definitions of orthometric heights. J Geod 80:691–704

    Article  Google Scholar 

  • Sheng MB, Shaw C, Vaníček P, Kingdon RW, Santos M, Foroughi I (2019) Formulation and validation of a global laterally varying topographical density model. Tectonophysics 762:45–60

    Article  Google Scholar 

  • Sjöberg LE (1995) On the quasigeoid to geoid separation. Manuscr Geod 20(3):182–192

    Google Scholar 

  • Sjöberg LE (1999) On the downward continuation error at the Earth’s surface and the geoid of satellite derived geopotential models. Boll Geod Sci Affini 58(3):215–229

    Google Scholar 

  • Sjöberg LE (2004) The effect on the geoid of lateral topographic density variations. J Geod 78:34–39

    Article  Google Scholar 

  • Sjöberg LE (2010) A strict formula for geoid-to-quasigeoid separation. J Geod 84:699–702

    Article  Google Scholar 

  • Sjöberg (2015) Rigorous geoid-from-quasigeoid correction using gravity disturbances. J Geod Sci 5:115–118

    Google Scholar 

  • Somigliana C (1929) Teoria Generale del Campo Gravitazionale dell’Ellisoide di Rotazione. Memoire della Societa Astronomica Italiana, IV, Milano

  • Stokes GG (1849) On the variation of gravity at the surface of the Earth. Trans Cambridge Philos Soc 9:672–695

    Google Scholar 

  • Stranska M, et al (1986) Hustotná mapa hornín Západných Karpát na území ČSSR. Final Report, MS Geofond, Bratislava

  • Tenzer R (2004) Discussion of mean gravity along the plumbline. Stud Geophys Geod 48:309–330

    Article  Google Scholar 

  • Tenzer R, Vaníček P, Santos M, Featherstone WE, Kuhn M (2005) The rigorous determination of orthometric heights. J Geod 79(1–3):82–92

    Article  Google Scholar 

  • Tenzer R, Moore P, Novák P, Kuhn M, Vaníček P (2006) Explicit formula for the geoid-to-quasigeoid separation. Stud Geoph Geod 50:607–618

    Article  Google Scholar 

  • Tenzer R, Sirguey P, Rattenbury M, Nicolson J (2011) A digital bedrock density map of New Zealand. Comput Geosci 37(8):1181–1191

    Article  Google Scholar 

  • Tenzer R, Novák P, Vajda P, Gladkikh V, Hamayun (2012a) Spectral harmonic analysis and synthesis of Earth’s crust gravity field. Comput Geosci 16(1):193–207

    Article  Google Scholar 

  • Tenzer R, Gladkikh V, Vajda P, Novák P (2012b) Spatial and spectral analysis of refined gravity data for modelling the crust-mantle interface and mantle-lithosphere structure. Surv Geophys 33(5):817–839

    Article  Google Scholar 

  • Tenzer R, Hirt Ch, Claessens S, Novák P (2015) Spatial and spectral representations of the geoid-to-quasigeoid correction. Surv Geophys 36(5):627–658

    Article  Google Scholar 

  • Tenzer R, Hirt Ch, Novák P, Pitoňák M, Šprlák M (2016) Contribution of mass density heterogeneities to the geoid-to-quasigeoid separation. J Geod 90(1):65–80

    Article  Google Scholar 

  • Tontini FC, Graziano F, Cocchi L, Carmisciano C, Stefanelli P (2007) Determining the optimal Bouguer density for a gravity data set: implications for the isostatic setting of the Mediterranean Sea. Geophys J Int 169(2):380–388

    Article  Google Scholar 

  • Toushmalani R, Saibi H (2015) 3D gravity inversion using Tikhonov regularization. Acta Geophys 63(4):1044–1065

    Article  Google Scholar 

  • Tziavos IN, Featherstone WE (2001) First results of using digital density data in gravimetric geoid computation in Australia. In: Sideris MG (ed) Gravity, Geoid and Geodynamics 2000. Springer, Berlin, pp 335–340

    Chapter  Google Scholar 

  • Vaníček P, Tenzer R, Sjöberg LE, Martinec Z, Featherstone WE (2005) New views of the spherical Bouguer gravity anomaly. Geophys J Int 159:460–472

    Article  Google Scholar 

  • Winester D (2016) USA National Surface Rock Density Map – Part 2. (Presented at AGU Fall Meeting, December 12 – 16, 2016. Poster: G11B-1080. San Francisco, USA)

  • Wirth B (1990) Höhensysteme, Schwerepotentiale und Niveauflächen. Geodätisch-Geophysikalische Arbeiten in der Schweiz, Vol 42, Swiss Geodetic Commission

  • Yang M, Hirt C, Tenzer R, Pail R (2018) Experiences with the use of mass-density maps in residual gravity forward modelling. Stud Geophys Geod 62:596–623

    Article  Google Scholar 

Download references

Acknowledgments

This research was conducted under the Hong Kong GRF science project Q73Z: The modernization of height datum in the Hong Kong territories.

Author information

Authors and Affiliations

Authors

Contributions

RT designed the study, derived corresponding formulas and drafted the manuscript. MP prepared the software. SR formatted input data and prepared figures. WC performed computations and numerical analysis. RT, WC, SR and MP read, commented and approved the final manuscript.

Corresponding author

Correspondence to Robert Tenzer.

Appendices

Appendix A: disturbing potential

The disturbing potential \( T \) (i.e., difference between values of the actual and normal gravity potential \( W \) and \( U \), respectively; \( T = W - U \)) at a location \( \left( {r,\varOmega } \right) \) is computed from (e.g., Heiskanen and Moritz 1967)

$$ T\left( {r,\varOmega } \right) = \frac{\text{GM}}{{\text{R}}}\;\sum\limits_{n = 0}^{{\bar{n}}} {\sum\limits_{m = - n}^{n} {\,\left( {\frac{{\text{R}}}{r}} \right)^{n + 1} \,{{\text{T}}}_{{\text{n,m}}} \,Y_{{\text{n,m}}} \left( \varOmega \right)} } , $$
(A.1)

where \( {\text{GM}} = 3 9 8 6 0 0 5\times 1 0^{ 8} \;{\text{m}}^{3} {\text{s}}^{ - 2} \) is the geocentric gravitational constant, \( {{\text{R}}} = 6 3 7 1\times 1 0^{ 3} \) m is the Earth’s mean radius, \( {{\text{T}}}_{{\text{n,m}}} \) are the disturbing potential coefficients, \( Y_{{\text{n,m}}} \) are the surface spherical functions of degree n and order m, and \( \bar{n} \) is the upper summation index of spherical harmonics. The 3D position in Eq. (A.1) and thereafter is defined in the spherical coordinate system \( \left( {r,\varOmega } \right) \), where \( r \) is a radius, and \( \varOmega = \left( {\varphi ,\lambda } \right) \) is a spherical direction with latitude \( \varphi \) and longitude \( \lambda \). The coefficients \( {{\text{T}}}_{{\text{n,m}}} \) are obtained from the Stokes coefficients of the global gravitational field after subtracting the coefficients describing the GRS80 (Moritz 2000) normal gravity field.

Appendix B: contribution of constant topographic density

From generalized expressions for the gravitational contribution of volumetric mass density contrast layer defined by Tenzer et al. (2015), the expression for computing the topographic potential difference for the constant topographic density in Eq. (7) is obtained in the following form (cf. Foroughi and Tenzer 2017)

$$ \begin{aligned} V_{g}^{{T,{{\rho }}^{{\text{T}}} }} - V_{t}^{{T,{{\rho }}^{{\text{T}}} }} & = \frac{\text{GM}}{{\text{R}}}\,\sum\limits_{n = 0}^{{\bar{n}}} {\left[ {1 - \left( {1 + \frac{H}{{\text{R}}}} \right)^{ - n - 1} } \right]} \\ &\quad \times \sum\limits_{m = - n}^{n} {{{\text{V}}}_{{\text{n,m}}}^{{{{\text{T}},\rho }^{{\text{T}}} }} \,{{\text{Y}}}_{{\text{n,m}}} \left( \varOmega \right)} \\ & \quad - \frac{\text{GM}}{{\text{R}}}\,\sum\limits_{n = 0}^{{\bar{n}}} {\sum\limits_{m = - n}^{n} {{{\text{V}}}_{{\text{n,m}}}^{\text{bias}} \,{{\text{Y}}}_{{\text{n,m}}} \left( \varOmega \right)} } . \\ \end{aligned} $$
(B.1)

The topographic \( {{\text{V}}}_{{\text{n,m}}}^{{{{\text{T}},\rho }^{{\text{T}}} }} \) and topographic-bias \( {{\text{V}}}_{{\text{n,m}}}^{{{{\text{T}},\rho }^{{\text{T}}} }} \) coefficients are defined by

$$ {{\text{V}}}_{{\text{n,m}}}^{{{{\text{T}},\rho }^{{\text{T}}} }} = \frac{3}{2n + 1}\;\frac{{{{\rho }}^{{\text{T}}} }}{{{\bar{{\rho }}}^{\text{Earth}} }}\sum\limits_{k = 0}^{n + 2} {\left( {\begin{array}{*{20}c} {n + 2} \\ k \\ \end{array} } \right)} \frac{1}{k + 1}\frac{{{{\text{H}}}_{{\text{n,m}}}^{{\left( {{\text{k}} + 1} \right)}} }}{{{{\text{R}}}^{k + 1} }}, $$
(B.2)
$$ {{\text{V}}}_{{\text{n,m}}}^{\text{bias}} = 3\frac{{{{\rho }}^{{\text{T}}} }}{{{\bar{{\rho }}}^{\text{Earth}} }}\sum\limits_{k = 1}^{2} {\frac{1}{k + 1}} \frac{{{{\text{H}}}_{{\text{n,m}}}^{{\left( {{\text{k}} + 1} \right)}} }}{{{{\text{R}}}^{k + 1} }}, $$
(B.3)

where \( \bar{\rho }^{\text{Earth}} = 5500 \) kg m−3 is the Earth’s mean density, and \( {{\rho }}^{{\text{T}}} \) is the average topographic density. The Laplace harmonics \( {{\text{H}}}_{\text{n}} \) of orthometric heights \( H \) (\( H \ge 0 \) on land and \( H = 0 \) offshore) are defined in terms of the following integral convolution (e.g., Sjöberg 1995)

$$ \begin{aligned} H\left( \varOmega \right) & = \sum\limits_{n = 0}^{\infty } {{{\text{H}}}_{\text{n}} \left( \varOmega \right)} ,\quad \\ {{\text{H}}}_{\text{n}} \left( \varOmega \right) & = \frac{2n + 1}{{4\uppi}}\iint\limits_{\varPhi } {H^{\prime}\,{\text{P}}_{\text{n}} \left( t \right)\;d}\varOmega^{\prime} = \sum\limits_{m = - n}^{n} {{{\text{H}}}_{{\text{n,m}}} \,{{\text{Y}}}_{{\text{n,m}}} \left( \varOmega \right)} , \\ \end{aligned} $$
(B.4)

where \( {{\text{H}}}_{{\text{n,m}}} \) are the height coefficients, \( H^{\prime} \) is the orthometric height, \( d\varOmega^{\prime} = \cos \varphi^{\prime}\,d\varphi^{\prime}\,d\lambda^{\prime} \) is the infinitesimal spherical surface element, and \( \varPhi = \left\{ {\;\varOmega^{\prime} = \left( {\varphi^{\prime},\lambda^{\prime}} \right):\varphi^{\prime} \in \left[ { - \pi /2,\,\pi /2} \right] \wedge \lambda^{\prime} \in \left[ {\left. {0,\,2\pi } \right)} \right.\;} \right\} \) is the full spatial angle. The Legendre polynomials \( {\text{P}}_{\text{n}} \) are defined for the argument \( t = \cos \psi \), where \( \psi \) is a spherical distance between points \( \left( {r,\varOmega } \right)\; \) and \( \left( {r^{\prime},\varOmega^{\prime}} \right) \). The corresponding higher-order harmonics { \( {{\text{H}}}_{\text{n}}^{{\left( {\text{k}} \right)}} :\;\,k = 2,\,3, \ldots \) } read (e.g., Tenzer et al. 2015)

$$ {{\text{H}}}_{\text{n}}^{{\left( {\text{k}} \right)}} \left( \varOmega \right) = \frac{2n + 1}{{4\uppi}}\iint\limits_{\varPhi } {H^{{\prime }{k}} \,{\text{P}}_{\text{n}} \left( t \right)\;d}\varOmega^{\prime} = \sum\limits_{m = - n}^{n} {{{\text{H}}}_{{\text{n,m}}} \,{{\text{Y}}}_{{\text{n,m}}} \left( \varOmega \right)} , $$
(B.5)

where the superscript k denotes the k-th integer power of orthometric heights.

Appendix C: contribution of anomalous lateral topographic density

The topographic potential difference \( V_{g}^{{T,\delta \rho^{T} }} - V_{t}^{{T,\delta \rho^{T} }} \) for the anomalous lateral topographic density (in Eq. 7) is computed from

$$ \begin{aligned} &\mathop { \lim }\limits_{{r \to {{{\rm R}}}^{ - } }} V_{\text{i}}^{{T,\delta \rho^{T} }} \left( {r,\varOmega } \right) - V_{{\rm e}}^{{T,\delta \rho^{T} }} \left( {r_{t} ,\varOmega } \right) \\ & \quad = \frac{\text{GM}}{{\text{R}}}\;\sum\limits_{n = 0}^{{\bar{n}}} {\,\sum\limits_{m = - n}^{n} {\,\left[ {{}_{{\rm i}}{{\text{V}}}_{{{\rm n,m}}}^{{T,\delta \rho^{T} }} - \left( {1 + \frac{H}{{\text{R}}}} \right)^{ - n - 1} {}_{{\rm e}}{{\text{V}}}_{{\rm n,m}}}^{{{T,\delta \rho^{T} }} } \right]\,Y_{{{\rm n,m}}} \left( \varOmega \right)} } . \\ \end{aligned} $$
(C.1)

The potential coefficients \( {}_{{\text{e}}}{{\text{V}}}_{{\text{n,m}}}^{{T,\delta \rho^{T} }} \) and \( {}_{\text{i}}{{\text{V}}}_{{\text{n,m}}}^{{T,\delta \rho^{T} }} \) of the anomalous topographic density contrast in Eq. (C.1) are given by

$$ {}_{{\text{e}}}V_{{\text{n,m}}}^{{T,\delta \rho^{T} }} = \frac{3}{2n + 1}\frac{{{}_{{\text{e}}}Fu_{{\text{n,m}}}^{{T,\delta \rho^{T} }} }}{{{\bar{{\rho }}}^{\text{Earth}} }}, $$
(C.2)
$$ {}_{\text{i}}V_{{\text{n,m}}}^{{T,\delta \rho^{T} }} = \frac{3}{2n + 1}\frac{{{}_{\text{i}}Fu_{{\text{n,m}}}^{{T,\delta \rho^{T} }} }}{{{\bar{{\rho }}}^{\text{Earth}} }}, $$
(C.3)

where the numerical coefficients \( {}_{{\text{e}}}Fu_{{\text{n,m}}}^{{T,\delta \rho^{T} }} \) and \( {}_{\text{i}}Fu_{{\text{n,m}}}^{{T,\delta \rho^{T} }} \) read

$$ {}_{{\text{e}}}Fu_{{\text{n,m}}}^{{T,\delta \rho^{T} }} = \sum\limits_{k = 0}^{n + 2} {\left( {\begin{array}{*{20}c} {n + 2} \\ k \\ \end{array} } \right)} \frac{1}{k + 1}\frac{{{{\delta \rho }}^{{\text{T}}} {\text{K}}_{{\text{n,m}}}^{{\left( {{\text{k}} + 1} \right)}} }}{{{{\text{R}}}^{{{\text{k}} + 1}} }}, $$
(C.4)
$$ {}_{\text{i}}Fu_{{\text{n,m}}}^{{T,\delta \rho^{T} }} = \sum\limits_{k = 0}^{\infty } {\left( { - 1} \right)^{k} \left( {\begin{array}{*{20}c} {n + k - 2} \\ k \\ \end{array} } \right)} \frac{1}{k + 1}\frac{{{{\delta \rho }}^{{\text{T}}} {\text{K}}_{{\text{n,m}}}^{{\left( {{\text{k}} + 1} \right)}} }}{{{{\text{R}}}^{{{\text{k}} + 1}} }}. $$
(C.5)

The combined density-heights coefficients \( \left\{ {{{\delta \rho }}^{{\text{T}}} {\text{K}}_{\text{n}}^{{\left( {\text{k}} \right)}} :\;\,k = 1,2,\,3, \ldots } \right\} \) in Eqs. (C4) and (C.5) are given by

$$ \begin{aligned} {{\delta \rho }}^{{\text{T}}} {\text{K}}_{\text{n}}^{{\left( {\text{k}} \right)}} \left( \varOmega \right) & = \frac{2n + 1}{{4\uppi}}\iint\limits_{\varPhi } {\delta \rho^{T} \left( {\varOmega^{\prime}} \right)K^{\prime k} \,{\text{P}}_{\text{n}} \left( t \right)\;{\text{d}}}\varOmega^{\prime} \\ & = \sum\limits_{m = - n}^{n} {{{\delta \rho }}^{{\text{T}}} {\text{K}}_{{\text{n,m}}} \,{{\text{Y}}}_{{\text{n,m}}} \left( \varOmega \right)} , \\ \end{aligned} $$
(C.6)

where the height \( K \) is equal to the orthometric height \( H \) except for areas covered by polar glaciers and lakes. In these two cases, the height \( K \) is defined as the orthometric height minus either the ice thickness or the inland bathymetric depth. The anomalous lateral topographic density \( \delta \rho^{T} \) in Eq. (C.6) is taken with respect to the constant topographic density \( {{\rho }}^{{\text{T}}} \). We then write

$$ \delta \rho^{T} \left( \varOmega \right) = {{\rho }}^{{\text{T}}} - \rho^{T} \left( \varOmega \right). $$
(C.7)

Appendix D: contribution of lakes and glaciers

The potential differences \( V_{g}^{L} - V_{t}^{L} \) and \( V_{g}^{I} - V_{t}^{I} \) in Eq. (7) of the gravitational contributions of lakes and polar glaciers are computed using the following generalized expression (cf. Foroughi and Tenzer 2017)

$$ \begin{aligned} & \mathop {\lim }\limits_{{r \to {{\text{R}}}^{ - } }} V_{\text{i}}^{\delta \rho } \left( {r,\varOmega } \right) - V_{{\text{e}}}^{\delta \rho } \left( {r_{t} ,\varOmega } \right) \\ & \quad = \frac{\text{GM}}{{\text{R}}}\;\sum\limits_{n = 0}^{{\bar{n}}} {\,\sum\limits_{m = - n}^{n} {\,\left[ {{}_{\text{i}}{{\text{V}}}_{{\text{n,m}}}^{{{{\delta \rho }}}} - \left( {1 + \frac{H}{{\text{R}}}} \right)^{ - n - 1} {}_{{\text{e}}}{{\text{V}}}_{{\text{n,m}}}^{{{{\delta \rho }}}} } \right]\,Y_{{\text{n,m}}} \left( \varOmega \right)} } . \\ \end{aligned} $$
(D.1)

The potential coefficients \( {}_{{\text{e}}}{{\text{V}}}_{{\text{n,m}}}^{{{{\delta \rho }}}} \) and \( {}_{\text{i}}{{\text{V}}}_{{\text{n,m}}}^{{{{\delta \rho }}}} \) in Eq. (D.1) read

$$ {}_{{\text{e}}}V_{{\text{n,m}}}^{{{{\delta \rho }}}} = \frac{3}{2n + 1}\frac{{{{\delta \rho }}}}{{{\bar{{\rho }}}^{\text{Earth}} }}\left( {{}_{{\text{e}}}Fu_{{\text{n,m}}}^{{}} - {}_{{\text{e}}}Fl_{{\text{n,m}}}^{{}} } \right), $$
(D.2)
$$ {}_{\text{i}}V_{{\text{n,m}}}^{{{{\delta \rho }}}} = \frac{3}{2n + 1}\frac{{{{\delta \rho }}}}{{{\bar{{\rho }}}^{\text{Earth}} }}\left( {{}_{\text{i}}Fu_{{\text{n,m}}}^{{}} - {}_{\text{i}}Fl_{{\text{n,m}}}^{{}} } \right), $$
(D.3)

where

$$ \begin{aligned} {}_{{\text{e}}}Fl_{{\text{n,m}}}^{{}} & = \sum\limits_{k = 0}^{n + 2} {\left( {\begin{array}{*{20}c} {n + 2} \\ k \\ \end{array} } \right)} \frac{1}{k + 1}\frac{{L_{{\text{n,m}}}^{{\left( {k + 1} \right)}} }}{{{{\text{R}}}^{{{\text{k}} + 1}} }},\quad \\ {}_{{\text{e}}}Fu_{{\text{n,m}}}^{{}} & = \sum\limits_{k = 0}^{n + 2} {\left( {\begin{array}{*{20}c} {n + 2} \\ k \\ \end{array} } \right)} \frac{1}{k + 1}\frac{{{{\text{H}}}_{{\text{n,m}}}^{{\left( {k + 1} \right)}} }}{{{{\text{R}}}^{{{\text{k}} + 1}} }}, \\ \end{aligned} $$
(D.4)
$$ \begin{aligned} {}_{\text{i}}Fl_{{\text{n,m}}}^{{}} & = \sum\limits_{k = 0}^{\infty } {\left( { - 1} \right)^{k} \left( {\begin{array}{*{20}c} {n + k - 2} \\ k \\ \end{array} } \right)} \frac{1}{k + 1}\frac{{L_{{\text{n,m}}}^{{\left( {k + 1} \right)}} }}{{{{\text{R}}}^{{{\text{k}} + 1}} }},\quad \\ {}_{\text{i}}Fu_{{\text{n,m}}}^{{}} & = \sum\limits_{k = 0}^{\infty } {\left( { - 1} \right)^{k} \left( {\begin{array}{*{20}c} {n + k - 2} \\ k \\ \end{array} } \right)} \frac{1}{k + 1}\frac{{{{\text{H}}}_{{\text{n,m}}}^{{\left( {k + 1} \right)}} }}{{{{\text{R}}}^{{{\text{k}} + 1}} }}. \\ \end{aligned} $$
(D.5)

The coefficients defined in Eqs. (D.4) and (D.5) utilize the spherical lower-bound functions \( {{\text{L}}}_{\text{n}} \) of a volumetric mass density contrast layer and their higher-order terms (cf. Tenzer et al. 2012a, 2012b)

$$ \begin{aligned} L_{\text{n}}^{{\left( {k + 1} \right)}} \left( \varOmega \right) & = \frac{2n + 1}{{4\uppi}}\iint\limits_{\varPhi } {H_{L}^{k + 1} \left( {\varOmega^{\prime}} \right)P_{n} \left( t \right)\;{\text{d}}\varOmega^{\prime}} \\ & = \sum\limits_{m = - n}^{n} {L_{{\text{n,m}}}^{{\left( {k + 1} \right)}} {{\text{Y}}}_{{\text{n,m}}} \left( \varOmega \right)} . \\ \end{aligned} $$
(D.6)

Since the upper bound of lakes and glaciers is identical with the topographic relief, the numerical coefficients \( {}_{{\text{e}}}Fu_{{\text{n,m}}}^{{}} \) and \( {}_{\text{i}}Fu_{{\text{n,m}}}^{{}} \) were generated directly from the height coefficients {\( {{\text{H}}}_{\text{n}}^{{\left( {\text{k}} \right)}} :\;\,k = 1,2, \ldots \) }; see Eqs. (B.4) and (B.5).

The density contrast \( {{\delta \rho }} \) is taken with respect to a constant topographic density \( {{\rho }}^{{\text{T}}} \). We then write

$$ {{\delta \rho }} = {{\rho }}^{{\text{T}}} - {{\rho }}\quad for\quad {{\text{R}}} + H_{U} \left( \varOmega \right) \ge r > {{\text{R}}} + H_{L} \left( \varOmega \right), $$
(D.7)

Appendix E: sub-geoid mass density contribution

To evaluate the no-topography disturbing potential difference \( T_{g}^{NT} - T_{t}^{NT} \) in Eq. (7), the complete topographic effect (including lakes and glaciers) is first subtracted from the disturbing potential \( T \) on the topographic surface. This procedure yields the no-topography disturbing potential \( T^{NT} \). The computation of \( T^{NT} \) is realized according to (Tenzer et al. 2015)

$$ T^{\text{n,m}} \left( {r_{t} ,\varOmega } \right) = \frac{\text{GM}}{{\text{R}}}\;\sum\limits_{n = 0}^{{\bar{n}}} {\sum\limits_{m = - n}^{n} {\,\left( {\frac{{\text{R}}}{{{{\text{R}}} + H}}} \right)^{n + 1} {{\text{T}}}_{{\text{n,m}}}^{\text{n,m}} Y_{{\text{n,m}}} \left( \varOmega \right)} } . $$
(E.1)

The (fully normalized) coefficients \( {{\text{T}}}_{{\text{n,m}}}^{\text{n,m}} \) are defined by

$$ {{\text{T}}}_{{\text{n,m}}}^{\text{n,m}} = {{\text{T}}}_{{\text{n,m}}} \, - {{\text{V}}}_{{\text{n,m}}}^{{{{\text{T}},\rho }^{{\text{T}}} }} + {}_{{\text{e}}}V_{{\text{n,m}}}^{{\text{L}}} + {}_{{\text{e}}}V_{{\text{n,m}}}^{\text{I}} + {}_{{\text{e}}}{{\text{V}}}_{{\text{n,m}}}^{{{\text{T,}}\delta \rho^{{\text{T}}} }} , $$
(E.2)

where \( {}_{{\text{e}}}V_{{\text{n,m}}}^{{\text{L}}} \), \( {}_{{\text{e}}}V_{{\text{n,m}}}^{\text{I}} \) and \( {}_{{\text{e}}}{{\text{V}}}_{{\text{n,m}}}^{{{\text{T,}}\delta \rho^{{\text{T}}} }} \) are, respectively, the coefficients of lakes, glaciers and anomalous topographic density.

The coefficients \( {{\text{T}}}_{{\text{n,m}}}^{\text{n,m}} \) are then used to compute the no-topography disturbing potential difference in Eq. (7) as follows (Tenzer et al. 2015)

$$ \begin{aligned} T_{g}^{NT} - T_{t}^{NT} & = T^{NT} \left( {r_{g} ,\varOmega } \right) - T^{NT} \left( {r_{t} ,\varOmega } \right) \\ & = \frac{\text{GM}}{{\text{R}}}\;\sum\limits_{n = 0}^{{\bar{n}}} \sum\limits_{m = - n}^{n} \left[ {\,1 - \left( {1 + \frac{H}{{\text{R}}}} \right)^{ - n - 1} } \right]\\ & \quad {{\text{T}}}_{{\text{n,m}}}^{\text{n,m}} \,Y_{{\text{n,m}}} \left( \varOmega \right) . \\ \end{aligned} $$
(E.3)

A downward continuation of \( T^{NT} \) in Eq. (E.3) is permissible, because topographic masses above the geoid are mathematically removed.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Tenzer, R., Chen, W., Rathnayake, S. et al. The effect of anomalous global lateral topographic density on the geoid-to-quasigeoid separation. J Geod 95, 12 (2021). https://doi.org/10.1007/s00190-020-01457-6

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1007/s00190-020-01457-6

Keywords

Navigation