Skip to main content

Advertisement

Log in

Gravity Spectra from the Density Distribution of Earth’s Uppermost 435 km

  • Published:
Surveys in Geophysics Aims and scope Submit manuscript

Abstract

The Earth masses reside in a near-hydrostatic equilibrium, while the deviations are, for example, manifested in the geoid, which is nowadays well determined by satellite gravimetry. Recent progress in estimating the density distribution of the Earth allows us to examine individual Earth layers and to directly see how the sum approaches the observed anomalous gravitational field. This study evaluates contributions from the crust and the upper mantle taken from the LITHO1.0 model and quantifies the gravitational spectra of the density structure to the depth of 435 km. This is done without isostatic adjustments to see what can be revealed with models like LITHO1.0 alone. At the resolution of 290 km (spherical harmonic degree 70), the crustal contribution starts to dominate over the upper mantle and at about 150 km (degree 130) the upper mantle contribution is nearly negligible. At the spatial resolution \(<150\,\hbox {km},\) the spectra behavior is driven by the crust, the mantle lid and the asthenosphere. The LITHO1.0 model was furthermore referenced by adding deeper Earth layers from ak135, and the gravity signal of the merged model was then compared with the observed satellite-only model GOCO05s. The largest differences are found over the tectonothermal cold and old (such as cratonic), and over warm and young areas (such as oceanic ridges). The misfit encountered comes from the mantle lid where a velocity–density relation helped to reduce the RMS error by 40%. Global residuals are also provided in terms of the gravitational gradients as they provide better spatial localization than gravity, and there is strong observational support from ESA’s satellite gradiometry mission GOCE down to the spatial resolution of 80–90 km.

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

Similar content being viewed by others

References

  • Afonso JC, Rawlinson N, Yang Y, Schutt DL, Jones AG, Fullea J, Griffin WL (2016) 3-D multiobservable probabilistic inversion for the compositional and thermal structure of the lithosphere and upper mantle: III. Thermochemical tomography in the Western-Central US. J Geophys Res Solid Earth 121(10):7337–7370

    Article  Google Scholar 

  • Anderson DL (2007) New theory of the earth. Cambridge University Press, Cambridge

    Book  Google Scholar 

  • Arfken GB, Weber HJ (2005) Mathematical methods for physicists international student edition. Academic press, Cambridge

    Google Scholar 

  • Artemieva IM (2006) Global \(1\times 1\) thermal model TC1 for the continental lithosphere: implications for lithosphere secular evolution. Tectonophysics 416(1):245–277

    Article  Google Scholar 

  • Auer L, Boschi L, Becker T, Nissen-Meyer T, Giardini D (2014) Savani: a variable resolution whole-mantle model of anisotropic shear velocity variations based on multiple data sets. J Geophys Res Solid Earth 119(4):3006–3034

    Article  Google Scholar 

  • Bertotti B, Farinella P, Vokrouhlicky D (2012) Physics of the solar system: dynamics and evolution, space physics, and spacetime structure, vol 293. Springer, Berlin

    Google Scholar 

  • Bird P (2003) An updated digital model of plate boundaries. Geochem Geophys Geosyst 4(3):

  • Bouman J, Ebbing J, Meekes S, Fattah RA, Fuchs M, Gradmann S, Haagmans R, Lieb V, Schmidt M, Dettmering D et al (2015) GOCE gravity gradient data for lithospheric modeling. Int J Appl Earth Obs Geoinf 35:16–30

    Article  Google Scholar 

  • Bouman J, Ebbing J, Fuchs M, Sebera J, Lieb V, Szwillus W, Haagmans R, Novak P (2016) Satellite gravity gradient grids for geophysics. Sci Rep. https://doi.org/10.1038/srep21050

    Google Scholar 

  • Chase CG (1979) Subduction, the geoid, and lower mantle convection. Nature 282:29

    Article  Google Scholar 

  • Cubells, J., Calsamiglia, A. (2010). Transitando por los espacios jurídico-penales: Discursos sociales e implicaciones para la intervención en casos de violencia hacia la mujer. Acciones e Investigaciones Sociales 28, 79-108

  • De Pater I, Lissauer JJ (2015) Planetary sciences. Cambridge University Press, Cambridge

    Book  Google Scholar 

  • Denis C, Rogister Y, Amalvict M, Delire C, Denis AI, Munhoven G (1997) Hydrostatic flattening, core structure, and translational mode of the inner core. Phys Earth Planet Inter 99(3):195–206

    Article  Google Scholar 

  • Ebbing J, Braitenberg C, Wienecke S (2007) Insights into the lithospheric structure and tectonic setting of the Barents Sea region from isostatic considerations. Geophys J Int 171(3):1390–1403. https://doi.org/10.1111/j.1365-246X.2007.03602.x

    Article  Google Scholar 

  • Ebbing J, Bouman J, Fuchs M, Lieb V, Haagmans R, Meekes J, Fattah RA (2013) Advancements in satellite gravity gradient data for crustal studies. Lead Edge 32(8):900–906

    Article  Google Scholar 

  • Floberghagen R, Fehringer M, Lamarre D, Muzi D, Frommknecht B, Steiger C, Piñeiro J, Da Costa A (2011) Mission design, operation and exploitation of the gravity field and steady-state ocean circulation explorer mission. J Geodesy 85(11):749–758

    Article  Google Scholar 

  • Fukao Y, Obayashi M (2013) Subducted slabs stagnant above, penetrating through, and trapped below the 660 km discontinuity. J Geophys Res Solid Earth 118(11):5920–5938

    Article  Google Scholar 

  • Fullea J, Rodríguez-González J, Charco M, Martinec Z, Negredo A, Villaseñor A (2015) Perturbing effects of sub-lithospheric mass anomalies in GOCE gravity gradient and other gravity data modelling: application to the Atlantic-Mediterranean transition zone. Int J Appl Earth Obs Geoinf 35:54–69

    Article  Google Scholar 

  • Gruber T (2015) GOCE gravity field models-signal and error assessment. In: EGU general assembly conference abstracts, vol 17, p 1657

  • Haagmans R (2000) A synthetic earth for use in geodesy. J Geodesy 74(7–8):503–511

    Article  Google Scholar 

  • Hager BH, Clayton RW, Richards MA, Comer RP, Dziewonski AM (1985) Lower mantle heterogeneity, dynamic topography and the geoid. Nature 313:541–545. https://doi.org/10.1038/313541a0

    Article  Google Scholar 

  • van Hees GS (2000) Some elementary relations between mass distributions inside the earth and the geoid and gravity field. J Geodyn 29(1):111–123

    Article  Google Scholar 

  • Hildebrand FB (1987) Introduction to numerical analysis. Courier Corporation, North Chelmsford

    Google Scholar 

  • James R, Kopal Z (1962) The equilibrium figures of the earth and the major planets. Icarus 1(1–6):442–454

    Article  Google Scholar 

  • Kaban M, Tesauro M, Cloetingh S (2010) An integrated gravity model for Europe’s crust and upper mantle. Earth Planet Sci Lett 296(3):195–209

    Article  Google Scholar 

  • Kaban MK, Schwintzer P, Artemieva IM, Mooney WD (2003) Density of the continental roots: compositional and thermal contributions. Earth Planet Sci Lett 209(1):53–69

    Article  Google Scholar 

  • Kaula WM (2000) Theory of satellite geodesy: applications of satellites to geodesy. Dover Publications, Mineola

    Google Scholar 

  • Kennett B, Engdahl E, Buland R (1995) Constraints on seismic velocities in the Earth from traveltimes. Geophys J Int 122(1):108–124

    Article  Google Scholar 

  • Laske G, Masters G, Ma Z, Pasyanos M (2013) Update on CRUST1. 0A 1-degree global model of Earth’s crust. Geophys Res Abstr 15:2658

    Google Scholar 

  • Li C, Van Der Hilst RD (2010) Structure of the upper mantle and transition zone beneath Southeast Asia from traveltime tomography. J Geophys Res Solid Earth 115(B7):

  • Martinec Z (2014) Mass-density Green’s functions for the gravitational gradient tensor at different heights. Geophys J Int 196(3):1455–1465

    Article  Google Scholar 

  • Mayer-Guerr T (2015) The combined satellite gravity field model GOCO05s. In: EGU general assembly conference abstracts, vol 17, p 12364

  • Montagner JP, Anderson DL (1989) Constrained reference mantle model. Phys Earth Planet Inter 58(2–3):205–227

    Article  Google Scholar 

  • Moritz H (2000) Geodetic reference system 1980. J Geodesy 74(1):128–133

    Article  Google Scholar 

  • Nolet G, Allen R, Zhao D (2007) Mantle plume tomography. Chem Geol 241(3):248–263

    Article  Google Scholar 

  • Panet I, Pajot-Métivier G, Greff-Lefftz M, Métivier L, Diament M, Mandea M (2014) Mapping the mass distribution of Earth’s mantle using satellite-derived gravity gradients. Nat Geosci 7(2):131–135

    Article  Google Scholar 

  • Pasyanos M (2017) Personal communication

  • Pasyanos ME, Masters TG, Laske G, Ma Z (2014) LITHO1.0: an updated crust and lithospheric model of the earth. J Geophys Res Solid Earth 119(3):2153–2173

    Article  Google Scholar 

  • Simmons NA, Forte AM, Boschi L, Grand SP (2010) GyPSuM: a joint tomographic model of mantle density and seismic wave speeds. J Geophys Res Solid Earth 115(B12):

  • Simmons NA, Myers SC, Johannesson G, Matzel E (2012) LLNL-G3Dv3: Global P wave tomography model for improved regional and teleseismic travel time prediction. J Geophys Res Solid Earth 117(B10):

  • Steinberger B, Becker TW (2016) A comparison of lithospheric thickness models. Tectonophysics

  • Tapley BD, Bettadpur S, Watkins M, Reigber C (2004) The gravity recovery and climate experiment: Mission overview and early results. Geophys Res Lett 31(9):

  • Tenzer R, Hamayun K, Vajda P (2009) Global maps of the crust 2.0 crustal components stripped gravity disturbances. J Geophys Res Solid Earth 114(B5):b05408. https://doi.org/10.1029/2008JB006016

    Article  Google Scholar 

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

    Article  Google Scholar 

  • Tenzer R, Chen W, Tsoulis D, Bagherbandi M, Sjöberg LE, Novák P, Jin S (2015) Analysis of the refined CRUST1.0 crustal model and its gravity field. Surv Geophys 36(1):139–165

    Article  Google Scholar 

  • Turcotte D, Schubert G (2002) Geodynamics, 2nd edn. Cambridge University Press, New York

    Book  Google Scholar 

  • van der Meijde M, Pail R, Bingham R, Floberghagen R (2015) GOCE data, models, and applications: a review. Int J Appl Earth Obs Geoinf 35:4–15

    Article  Google Scholar 

  • Vallado D, McClain W (2001) Fundamentals of astrodynamics and applications, space technology library. Kluwer Academic Publishers, Dordrecht, p 792

    Google Scholar 

  • Yegorova T, Pavlenkova G (2015) Velocity-density models of the earth’s crust and upper mantle from the quartz, craton, and kimberlite superlong seismic profiles. Izvestiya Phys Solid Earth 51(2):250

    Article  Google Scholar 

  • Zoback ML, Mooney WD (2003) Lithospheric buoyancy and continental intraplate stresses. Int Geol Rev 45(2):95–118

    Article  Google Scholar 

Download references

Acknowledgements

The study is connected to the ESA STSE project “3D Earth - A Dynamic Living Planet” (https://www.3dearth.uni-kiel.de/en). We thank the Editor in Chief Michael J. Rycroft and anonymous reviewers for their helpful comments.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Josef Sebera.

Appendices

Appendix 1: Choice of Volume Elements

Triangular parameterizations such as that used in LITHO1.0 are not often used for calculating the Earth’s gravitational signal from a global density distribution; see, however, the exception (Haagmans 2000). The situation is complicated by the fact that each node might be equipped with a unique value of the density and individual triangles may differ in area. This causes an ambiguity in a way the volume element can be set up.

To reduce the effect of triangulation in LITHO1.0, we define the volume element of each node according to Fig. 7 as space between two spherical surfaces (\(R_1\) and \(R_2\)) for a specific solid angle:

$$\begin{aligned} {\hbox{d}}V' = \frac{1}{3} S_a(R_2^3 - R_1^3), \end{aligned}$$
(8)

where \(R_1<=R_2\) and \(S_a\) is the average solid angle from the all spherical triangles containing the particular node. The gray area on the left of Fig. 7 depicts \(S_a\), white radii bound the volume element vertically and the green area shows the discretized layer passing from node to node. Because the number of nodes is half of the number of triangles (40,962 nodes vs. 81,920 triangles), the local average \(S_a\) must be multiplied by two (or exactly by 81,920/40,962) to cover \(4\pi\). Nonetheless, a global sum of such average \(S_a\) will deviate from \(4\pi\). In our case, the error is \(\delta =4.4\cdot 10^{-5}\) steradian that corresponds to more than 5 significant decimal digits of \(4\pi\). To compare, a sum of all spherical triangles provided with LITHO1.0 gives \(4\pi\) with the error \(\delta =-3.5\cdot 10^{-10}\) steradian. Hence, the volume numerical integration with models like LITHO1.0 is a sort of a trade-off between a way the local average is made (to assign the density to a volume) and preservation of the whole Earth surface/mass.

The variation of the average surface elements in percent is shown on the right of Fig. 7. The patterns agree well with the process of tesselation used in LITHO1.0; nearly each tesselation level can be recognized (from large to small patterns). The area variation is computed with respect to the largest value (red) and it reaches up to 22%. For example, omitting this area variation and using a constant value for each node (i.e., \(S_a=4\pi /40962\)) affects all significant digits of the resulting gravitational signal.

Fig. 7
figure 7

On the left, the average solid angle of each node for setting up volume elements according to Eq. (8). On the right, the global variation of \(S_a\) in percent (relatively to the largest value) obtained from LITHO1.0

Appendix 2: Comparison of 1D and 3D Integration

One dimensional approximations such Eq. (7) are often being used to calculate the gravity signal from a density column \(\rho _i\) (Turcotte and Schubert 2002; Hees 2000). Although not by definition, the shell formula may also provide a quasi-lateral variation if each density column is evaluated separately. Hence, by using all the non-homogeneous layers in LITHO1.0 a 3D approach based on Eq. (4) can be confronted with Eq. (7).

Figure 8 shows such a difference in terms of \(g_U\). It is seen the large discrepancies are located over the thick parts of the lithosphere and over the oceanic ridges where the shell formula simply does not account for a lateral density variation. The magnitudes reach \(\pm 1000\,\hbox {mGal}\) that is more than twice as much as the difference between GOCO05s and LITHO1.0 (including ak135) evaluated with Eq. (2); see Fig. 5 and Table 1. The 1D approximation would introduce a large methodological error in our context. On the other hand, Eq. (7) is not singular (compare with Eq. (2) when \(L\rightarrow 0\)) and it is much faster to evaluate than Eq. 2 (seconds vs. hours on a single PC).

Fig. 8
figure 8

A difference in \(g_U\) (in mGal) between Eqs. (2) and  (7) at 6621 km by using all the non-homogeneous layers in LITHO1.0

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Sebera, J., Haagmans, R., Floberghagen, R. et al. Gravity Spectra from the Density Distribution of Earth’s Uppermost 435 km. Surv Geophys 39, 227–244 (2018). https://doi.org/10.1007/s10712-017-9445-z

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s10712-017-9445-z

Keywords

Navigation