Introduction

More than 150 INTERMAGNET geomagnetic observatories around the world readily provide high-quality magnetic field data with a sampling rate of up to 1 Hz (St-Louis 2012). In principle, the recorded field variations with periods up to a few hours would allow to resolve the electrical conductivity distribution in the Earth’s lithosphere and upper mantle. However, there are two limitations impeding such studies. First, geomagnetic observatories are sparse, hence for studies of the lithosphere and upper mantle only individual observatories can be considered. In consequence, one can at best resolve a one-dimensional (1-D) conductivity structure beneath specific locations. Second, the source of field variations under consideration is seen at mid-latitude observatories as a time-varying vertically incident plane wave (Chave and Jones 2012). With such a source field, tippers are the only response functions that might be estimated from single observatories recording only geomagnetic data. However, tippers are zero for a purely 1-D conductivity distribution (Simpson and Bahr 2005; Berdichevsky and Dmitriev 2008), and therefore do not allow a meaningful interpretation with a single station.

Still, the situation is different for island and coastal observatories, where lateral conductivity contrasts between conductive sea water and more resistive island rock result in lateral conductivity variations and therefore large tipper amplitudes even if the underlying structure is essentially 1-D (Samrock and Kuvshinov 2013). This is referred to as the ocean (or coast) induction effect (Parkinson and Jones 1979). Berdichevsky and Dmitriev (2008) argued that tippers are also sensitive to vertical variations in conductivity. This was confirmed by Samrock and Kuvshinov (2013) using realistic three-dimensional (3-D) bathymetry and 1-D conductivity profiles that significantly differ within the lithosphere and upper mantle (\(< 200\) km). In their model study, the resulting variability in the modeled tippers exceeded the uncertainties in the observed tippers.

In this paper, we go a step further and develop a methodology to invert island tippers for a 1-D conductivity distribution in the presence of known 3-D bathymetry. Using this methodology, we first perform realistic synthetic tests and demonstrate that tippers from a single island observatory are indeed able to recover the synthetic 1-D conductivity profile. Then, we invert tippers obtained at two island observatories located in different tectonic environments, namely Gan (GAN) in the Indian Ocean and Tristan da Cunha (TDC) in the South Atlantic. Finally, we interpret the recovered 1-D profiles in terms of their regional geology and discuss the caveats of the 1-D assumption at remote oceanic islands.

Methods

Tippers

Tippers \(\mathbf {T}\) are the only single-site geomagnetic response functions within the plane-wave paradigm. They relate the horizontal (\(B_x\), \(B_y\)) and the vertical (\(B_z\)) magnetic field components as

$$\begin{aligned} B_z(\omega )= \begin{pmatrix} T_{zx}(\omega ) \; T_{zy}(\omega ) \end{pmatrix} \begin{pmatrix} B_{x}(\omega ) \\ B_{y}(\omega ) \end{pmatrix}. \end{aligned}$$
(1)

Here \(\omega = 2 \pi /p\) is the angular frequency of geomagnetic field variations, where p is a period, z points positive downward, and x and y point to geographic North and East, respectively. As a consequence of the plane-wave excitation, variations in \(B_z\) (and thus \(\mathbf {T}\)) are zero above 1-D conductivity structures (Simpson and Bahr 2005; Berdichevsky and Dmitriev 2008). Tippers are often represented in terms of real and imaginary induction arrows. In what follows, real induction arrows point away from conductive zones (Wiese convention, Wiese 1965) and are given by \((\mathfrak {R}{T_x}, \mathfrak {R}{T_y})\). We estimated tippers from time-series of geomagnetic observatory data using robust processing methods (Appendix A). Additionally, univariate coherencies as well as multivariate coherencies were calculated, i.e., \({\mathrm{coh}}(T_{zx})={\mathrm{corr}}(B_z,T_{zx}B_x)\) and \({\mathrm{coh}}(\mathbf {T})={\mathrm{corr}}(B_z,T_{zx}B_x+T_{zy}B_y)\), respectively. Here, \({\mathrm{corr}}\) refers to the Pearson correlation coefficient. In consequence, and as \(B_z\) is a linear combination of \(B_x\) and \(B_y\) (Eq. 1), univariate coherencies \({\mathrm{coh}}(T_{zx})\) and \({\mathrm{coh}}(T_{zy})\) cannot simultaneously have values close to one. Note that this is also true for any univariate coherencies contributing to the same output variable (such as, for example, \({\mathrm{coh}}(Z_{xx})\) and \({\mathrm{coh}}(Z_{xy})\) for impedances).

Numerical model

As is shown in Fig. 1a, we parameterized the model by dividing the subsurface into M “layers” with unknown homogeneous conductivities \(\sigma _i\). Depending on bathymetry, these layers may not extend throughout the modeling domain and can be interrupted by regions of seawater conductivity. Further, conductivities of the air and sea remained fixed, and the sea conductivity values were obtained from the World Ocean Atlas (available as supplementary material in Grayver et al. 2016).

Fig. 1
figure 1

Short title: Model parameterization Detailed legend: a Sketch of the adopted model parameterization. The \(\sigma _1,\ldots ,\sigma _M\) conductivity values denote the unknowns, which are estimated using data from the observatory. Note that the bottom layer extends to infinity. b Example of a locally refined mesh for the Tristan da Cunha island with topography-conforming mesh

The conductivity contrast between rocks and saline seawater results in a complex 3-D conductivity structure. In order to model it accurately, a 3-D forward solver with adaptive mesh refinement is used that solves Maxwell’s equations. The numerical mesh used for Tristan da Cunha is shown in Fig. 1b, and more details on the numerical scheme and solver can be found in Grayver and Kolev (2015) and in Appendix B.

Synthetic test

Samrock and Kuvshinov (2013) demonstrated that tippers are sensitive to different 1-D conductivity profiles in the presence of the ocean induction effect. Here, we perform synthetic tests to show that these differences are sufficient to infer the conductivity distribution with depth. For this purpose, we use the same bathymetry and numerical setup as for the real data from TDC (c.f. "Inversion of TDC data" section) and evaluate tippers at the location of the TDC observatory (see red diamond on inlet of Fig. 3b) for different profiles of electrical conductivity.

Fig. 2
figure 2

Short title: Synthetic test Detailed legend: a Conductivity profiles and b corresponding tippers for TDC bathymetry are shown for two synthetic conductivity profiles: a homogeneous Earth model [(HOM, dashed gray line (a) and dashed colored lines (b)], and a profile varying with depth (1-D, red line (a) and symbols (b)). The synthetic tippers that correspond to the 1-D profile were inverted with 3% added noise and resulted in a the solid black conductivity profile and b the tippers shown by solid colored lines

Fig. 3
figure 3

Short title: Bathymetry with response functions Detailed legend: The bathymetry (GEBCO30) for the GAN (a) and TDC (b) observatories is shown for the area used for numerical modeling, and the location of the observatories is indicated by a red diamond in the inlets. Please note that the Addu Atoll is barely above sea level, and green points mark these regions whereas the black line also includes regions down to 30 m below sea level. In the lower panels, the observed induction arrows, tippers, and coherencies are shown for the GAN (c) and TDC (d) observatories. As expected, the real (blue) induction arrows point toward the more resistive islands

First, we illustrate the significance of the ocean induction effect by calculating tippers for a model with ocean plus a homogeneous Earth model (HOM) of \(\rho =100\) \(\Omega {\text{m}}\) (dashed gray line in Fig. 2a). Without ocean induction effect, such a model would result in zero tippers. However, modeled tippers are significant as shown by dashed blue and red lines in Fig. 2b. Next, we choose a 1-D conductivity profile that is varying with depth, as indicated by the red line in Fig. 2a. The crosses in Fig. 2b show the resulting tippers which differ significantly from those corresponding to the HOM. This result confirms that tippers are sensitive to the underlying 1-D conductivity structure, as has been shown by Samrock and Kuvshinov (2013). Next, we add 3 % Gaussian noise to the synthetic tippers, as indicated by the error bars, and invert them for a 1-D conductivity profile. The synthetic tippers can be well fit (crosses and solid lines in Fig. 2b), and the obtained conductivity profile (black solid line in Fig. 2a) matches the true synthetic profile (red line) within its error limits (black dashed lines) as obtained from the diagonals of the model covariance matrix (i.e., only uncorrelated errors).

Observed tippers

Gan observatory (GAN)

The GAN geomagnetic observatory (Velimsky et al. 2014) is located on the Addu Atoll at the southern end of the Maldives archipelago, at \(0.6946^\circ\)S and \(73.1537^\circ\)E (red diamond in Fig. 3a). The Maldives archipelago consists of corals on top of the Maldive Ridge which was formed by the Réunion hot spot about 50–55 Ma ago (Fontaine et al. 2015, Fig. 1).

The top panel of Fig. 3c displays the observed induction arrows and the middle panel shows the real and imaginary parts of the corresponding tipper values. As expected, the real (blue) arrows point toward the more resistive island chain of the Maldives in the North (cf. Fig. 3a). Further, univariate coherencies (bottom panel) are significantly lower for \({\mathrm{coh}}(T_{zy})\) than for \({\mathrm{coh}}(T_{zx})\). This is due to the fact that both \({\mathrm{coh}}(T_{zy})\) and \({\mathrm{coh}}(T_{zx})\) cannot be simultaneously large ("Tippers" section), and that tippers are mainly oriented along the S–N axis which results in a better signal-to-noise ratio for \(T_{zx}\). Multivariate coherencies reach values \(> 0.8\) (black line), except for periods \(< 50\, {\text{s}}\) where the sensitivity of the fluxgate magnetometer starts to decrease.

Tristan da Cunha observatory (TDC)

The TDC magnetic observatory (Matzka et al. 2009, 2010) is located on Tristan da Cunha island at \(37.067^\circ\)S and \(12.315^\circ\)W (red diamond on inlet of Fig. 3b) with seafloor ages of 20–25 Myr (Müller et al. 2008). Tristan da Cunha is an active volcano in the South Atlantic Ocean, located at the western end of the Walvis Ridge that marks the Tristan-Gough hotspot track (Rohde et al. 2013).

The observed induction arrows, real and imaginary values of tippers, and coherencies are shown in Fig. 3d. As expected, the real arrows point toward the more resistive island of Tristan in the south-southeast (Fig. 3b), at least for periods up to  500 s. For longer periods, they tend to be influenced by the presence of the islands Nightingale and Inaccessible to the south-southwest (Fig. 3b). Again, univariate coherencies are lower for \({\mathrm{coh}}(T_{zy})\) than for \({\mathrm{coh}}(T_{zx})\), especially at longer periods where tippers are almost perfectly aligned with the S–N axis. Multivariate coherencies for tippers at TDC have values \(> 0.8\), except for periods \(< 20\, {\text{s}}\). Compared to GAN, multivariate coherencies at short periods are higher as more data were available.

Results and interpretation

Inversion of GAN data

The modeling domain covers an area of \(936 \times 936\) km with the GAN observatory at the center (Fig. 3a). Tippers were inverted for periods ranging between 40 and \(10{,}000\,\hbox {s}\), and a residual weighted RMS of 2.2 was achieved. The resulting best-fit tippers agree well with observed tippers, except for the real part of \(T_{zy}\) (Fig. 4b). The poor fit to \(T_{zy}\) may be related to the low coherency of \(T_{zy}\) (c.f. Fig. 3c). Indeed, additional induction coil data, shown by crosses in Fig. 4b, show a better agreement with modeled tippers (solid lines). Alternatively, the misfit may result from 3-D conductivity structures that are incompatible with the assumed 1-D model, for example conductive seafloor sediments at some distance from the station, or from inaccuracies of the bathymetry model. Disregarding \(T_{zy}\) slightly improves the fit to \(T_{zx}\) without corrupting the fit to \(T_{zy}\), as shown in Fig. 4b by comparing the observed (open circles) and modeled (solid lines) tippers.

Fig. 4
figure 4

Short title: 1-D conductivity profiles Detailed legend: Left column (a+c): the best-fit conductivity profiles derived using tippers at the GAN observatory (a) and the TDC observatory (c) are shown by black solid lines along with 95 % confidence intervals (black dashed lines). For reference, the conductivity profiles for old Pacific lithosphere (red line, Baba et al. 2010), for Tristan da Cunha (turqouise line, Baba et al. 2016), and hydrous and dry olivine (blue and brown lines, Katsura and Yoshino 2015) are shown. Additionally, best-fit homogeneous Earth models are shown by dashed gray lines. Right column (b+d): the observed tippers are indicated by circles, the tippers predicted from the 1-D models are indicated by solid lines, and the tippers predicted from the homogeneous Earth models are indicated by dashed lines. Additionally, induction coil data for GAN are shown by crosses

We validate the robustness of our model by additionally inverting \(T_{zx}\) for a homogeneous Earth model (HOM), resulting in an electrical conductivity of \(\sigma =24\) mS/m, and we note that an inversion to all the data results in a very similar conductivity of \(\sigma =25\) mS/m. The model is shown by the gray dashed line in Fig. 4a and resembles the conductivity of the 1-D model at depth. The corresponding tippers are shown by the dashed lines in Fig. 4b and fit the data with an RMS of 3.8 as compared to an RMS of 2.2 for the 1-D model. Clearly, the HOM Earth model is not able to reproduce all of the data. Therefore, additional variations in conductivity are required by the data, and here we assume that these additional variations are only depth dependent.

The conductivity model resulting from a fit to \(T_{zx}\) only (black line in Fig. 4a) is characterized by a thin layer of conductive sediments, a thick layer of resistive crust and lithosphere, and a subsequent increase in conductivity with depth. Compared to the model resulting from a fit to all data, it is only slightly more resistive in the crust and lithosphere (not shown). Along with the model, the resulting 95% confidence limits are provided, and we note that these depend on the data errors which in turn depend on the assumed error floor (see Appendix B).

The upper resistive layer probably characterizes volcanic and basement rock (Aubert and Droxler 1996), and the obtained upper mantle conductivity at depths \(>110\) km agrees with the conductivity of dry olivine (brown line in Fig. 4a) as calculated for a normal mantle geotherm with \(T = 1360^{\circ }\) C at the lithosphere-asthenosphere boundary (Katsura and Yoshino 2015). For comparison, the corresponding conductivity of wet olivine is shown by the blue line (Katsura and Yoshino 2015). Overall, the recovered 1-D conductivity profile resembles the conductivity of resistive 125–150 Myr old Pacific upper mantle (Baba et al. (2010), red line in Fig. 4a), and differences at depths of 110–170 km possibly result from lower sensitivity at these depths and smoothing regularization that penalizes any structure. In conclusion, our model suggests no thermal or compositional anomalies at the base of the lithosphere of the Maldives Ridge, in agreement with seismic tomography data (Fontaine et al. 2015).

Inversion of TDC data

TDC has less prominent regional bathymetry as compared to GAN (Fig. 3b), and therefore a numerical domain of \(586 \times 586\) km was sufficient. Further, tippers were modeled for periods between 60 and 8264 s, and an overall residual weighted RMS of 0.5 was achieved. The resulting conductivity profile is shown along with its confidence limits by the black solid and dashed lines in Fig. 4c, and the corresponding modeled and observed tippers are shown by the solid line and symbols in Fig. 4d. The model fits the data very well, and only the longest periods of \(Re(T_{zy})\) are not entirely fit within their error limits.

Similar as for GAN, we validate the robustness of our model by additionally inverting for a homogeneous Earth model (HOM). The obtained electrical conductivity of \(\sigma =20\) mS/m roughly corresponds to the average conductivity of the 1-D model. The corresponding tippers fit the data with an RMS of 1.1 as compared to an RMS of 0.5 for the 1-D model and are shown by the dashed lines in Fig. 4d. As for GAN, the HOM Earth model is not able to fit all of the data, especially for \(T_{zx}\).

Compared to GAN, conductivities are overall higher for TDC with a thinner resistive layer followed by a increase in conductivity that starts at \(\approx 30\) km depth. As for GAN, the conductivity of hydrous olivine (Katsura and Yoshino 2015) is shown by the blue line in Fig. 4c. In comparison with the obtained conductivities and their confidence interval, the presence of melt is therefore not required at depth, although it cannot be excluded either. In addition, the 3-D envelope of a recent 3-D MT sea-bottom survey by Baba et al. (2016) is shown by the turquoise lines in Fig. 4c. Baba et al. (2016) concluded that no plume-like structure below the island is necessary to explain the data, but the authors note that their survey layout would not sense small-scale (\(<150\) km) structures directly below the island due to the absence of measurements in the island’s vicinity. Other studies found indications for melt below TDC. For example, Weit et al. (2017) concluded from geochemical considerations that melt fractions of 5% are present at depths of 60–100 km. Also, seismic tomography suggests a crustal low-velocity anomaly characteristic for a magma feeding system below the island (Ryberg et al. 2017).

Overall, the electrical conductivities obtained by Baba et al. (2016) agree well with this study, with the exception of a more resistive crust and lithosphere, especially at depths of 20–50 km. This confirms the validity of our approach, and the potential to resolve 1-D conductivity profiles using tippers and geomagnetic observatory data. Still, we note that results should not be overinterpreted, as the presented approach is only useful if the 1-D conductivity assumption is not strongly violated, and if the used model of bathymetry is of good quality and resolution.

Discussion and conclusions

To our knowledge, tippers estimated at geomagnetic observatories have not been used for studying 1-D electrical conductivity distributions. However, tippers dominated by 3-D effects such as bathymetry also contain information on the vertical conductivity profile (Berdichevsky and Dmitriev 2008; Samrock and Kuvshinov 2013). In consequence, if the 3-D effects are known, the 1-D conductivity profile can be extracted.

Here, we demonstrated with synthetic tests that tippers can be used to resolve the 1-D conductivity structure below islands where 3-D effects are constrained by bathymetry. Further, we processed geomagnetic data from two island observatories, Gan in the Indian Ocean and Tristan da Cunha in the South Atlantic, to estimate and invert tippers. The obtained electrical conductivity profile beneath Gan island exhibits no anomalous behavior and is in agreement with other results for old oceanic upper mantle. For TDC, electrical conductivities are higher than for GAN, which may be explained by increased temperatures or increased water content. The presence of melt below the island is not required by the data, but cannot be excluded either. In the same area, an extensive 3-D MT sea-bottom survey (Baba et al. 2016) obtained electrical conductivities that are very similar to our results, confirming the applicability of the proposed method. Additionally, we test the robustness of the obtained 1-D profiles by inverting for a homogeneous Earth model and conclude that the additional 1-D variability is required by the data.

We note that our approach relies on the assumption that no additional 3-D conductivity anomalies are present. In consequence, any such anomalies may lead to misleading results, or to the unability of the model to fit the data. Clearly, remote and active volcanic islands are anomalous regions, for example due to hotspot mantle plumes. Still, the developed methodology allows us to obtain knowledge of the oceanic lithosphere and upper mantle from available single-site geomagnetic observatory data that otherwise cannot be used for induction studies. Therefore, as long as we are aware of its limitations, the method will help to better understand the oceanic lithosphere and upper mantle in remote regions where otherwise little or no knowledge is available. Furthermore, when combined with longer period responses due to Sq or magnetospheric sources, this methodology provides a unique opportunity for imaging the electrical structure of the mantle throughout its full depth range. In future, more complete 3-D studies with multiple magnetometers or additional measurements of field intensity (Kuvshinov et al. 2016) are also feasible.