Location via proxy:   [ UP ]  
[Report a bug]   [Manage cookies]                
License: CC BY-NC-ND 4.0
arXiv:2312.09878v1 [physics.geo-ph] 15 Dec 2023
\pagerange

Assimilation of ground and satellite magnetic measurements: inference of core surface magnetic and velocity field changesReferences

Assimilation of ground and satellite magnetic measurements: inference of core surface magnetic and velocity field changes

O. Barrois11{}^{1}start_FLOATSUPERSCRIPT 1 end_FLOATSUPERSCRIPT    M.D. Hammer22{}^{2}start_FLOATSUPERSCRIPT 2 end_FLOATSUPERSCRIPT    C.C. Finlay22{}^{2}start_FLOATSUPERSCRIPT 2 end_FLOATSUPERSCRIPT    Y. Martin11{}^{1}start_FLOATSUPERSCRIPT 1 end_FLOATSUPERSCRIPT & N. Gillet11{}^{1}start_FLOATSUPERSCRIPT 1 end_FLOATSUPERSCRIPT
11{}^{1}start_FLOATSUPERSCRIPT 1 end_FLOATSUPERSCRIPT Univ. Grenoble Alpes
   CNRS    ISTerre    CS 40700 F-38058 Grenoble cedex 9    France.
22{}^{2}start_FLOATSUPERSCRIPT 2 end_FLOATSUPERSCRIPT Division of Geomagnetism
   DTU Space    Technical University of Denmark    Lyngby DK-2800    Denmark
(Received XXX; in original form XXX; 2018)
keywords:
data assimilation – core dynamics – geomagnetic field modelling – satellite observations
volume: XXX{summary}

We jointly invert for magnetic and velocity fields at the core surface over the period 1997–2017, directly using ground-based observatory time series and measurements from the CHAMP and Swarm satellites. Satellite data are reduced to the form of virtual observatory time series distributed on a regular grid in space. Such a sequential storage helps incorporate voluminous modern magnetic data into a stochastic Kalman filter, whereby spatial constraints are incorporated based on a norm derived from statistics of a numerical geodynamo model. Our algorithm produces consistent solutions both in terms of the misfit to the data and the estimated posterior model uncertainties. We retrieve core flow features previously documented from the analysis of spherical harmonic field models, such as the eccentric anti-cyclonic gyre. We find enhanced diffusion patterns under both Indonesia and Africa. In contrast to a steady flow that is strong under the Atlantic hemisphere but very weak below the Pacific, interannual motions appear evenly distributed over the two hemispheres. Recovered interannual to decadal flow changes are predominantly symmetrical with respect to the equator outside the tangent cylinder. In contrast, under the Northern Pacific we find an intensification of a high latitude jet, but see no evidence for a corresponding feature in the Southern hemisphere. The largest flow accelerations that we isolate over the studied era are associated with meanders, attached to the equatorward meridional branch of the planetary gyre in the Eastern hemisphere, that is linked to the appearance of an eastward equatorial jet below the Western Pacific.

1 Introduction

Inferring information on the motions of the liquid outer core of the Earth requires properly separating the numerous sources of observed magnetic fields (geodynamo, crustal magnetisation, ionospheric and magnetospheric currents and their Earth induced counterparts). To circumvent some of the leakage issues, magnetic field models are often built using regularizations, to ensure spectral convergence of the core field and its time variations. This prevents a proper assessment of a posteriori errors on model coefficients. When these are used as data in reconstructions of the core dynamics, it can lead to biased estimates. Furthermore, by proceeding in successive steps (to a field model and then on to the core flow), one loses information.

From the early 1990’s alternative avenues of research arose, through which field models were built under topological constraints derived from physical insights. Constable et al.(1993)Constable, Parker, & Stark and O’Brien et al.(1997)O’Brien, Constable, & Parker proposed algorithms to apply, on single epoch pairs of models, magnetic flux conservation conditions at the core-mantle boundary (CMB) that are appropriate assuming that magnetic diffusion is negligible. Along the same lines, Jackson et al.(2007)Jackson, Constable, Walker, & Parker added a constraint on the radial vorticity. They showed that it was possible for a magnetic model to satisfy both these topological conditions, and the constraint from magnetic observations, from the late 19th century onwards.

Conversely, Chulliat & Olsen(2010) tested the validity of the frozen flux hypothesis using data from Magsat, Oersted and CHAMP satellite missions. They found an increase of the data misfit in some areas, potentially suggesting local failures of the constraint. Such studies motivated the co-estimation, from magnetic observations, of both the field and the flow, imposing with a weak formalism the frozen flux radial induction equation at the CMB (Lesur et al.(2010)Lesur, Wardinski, Asari, Minchev, & Mandea; Wardinski & Lesur(2012)). They concluded that the frozen flux constraint remained compatible with ground-based and satellite magnetic records. Pursuing an alternative approach, Beggan & Whaler(2009) and Whaler & Beggan(2015) obtained piecewise constant or linear flow models directly from magnetic data (see also Whaler et al.(2016)Whaler, Olsen, & Finlay).

One limitation though of such approaches is related to the uncertainties associated with the large scale induction equation itself (and associated null-flux curves), assuming models truncated at spherical harmonic degree n13similar-to-or-equals𝑛13n\simeq 13italic_n ≃ 13 (Gillet et al.(2009)Gillet, Pais, & Jault). Subgrid-scale effects arising due to the non-linear induction process (e.g. Eymin & Hulot(2005); Pais & Jault(2008); Gillet et al.(2009)Gillet, Pais, & Jault; Baerenzung et al.(2016)Baerenzung, Holschneider, & Lesur) turn out to be the main source of uncertainty in the recovery of core surface flows from modern geomagnetic records. Barrois et al.(2017)Barrois, Gillet, & Aubert – hereafter referred as BGA17 – illustrate how ignoring their impact leads to severely biased flow models (see also Baerenzung et al.(2017)Baerenzung, Holschneider, Wicht, Sanchez, & Lesur, on the reliability of core flow reconstructions).

BGA17 furthermore show from the analysis of geodynamo simulations that magnetic diffusion at the core surface, enslaved to poloidal flow below the CMB, affects the recorded field changes at all time-scales including rapid changes. This may seem at odds with the often used assumption of negligible magnetic diffusion that follows the argument of a high magnetic Reynolds number for large-scale motions in the core (see Holme(2015)).

In the present work we invert, from magnetic field observations collected at and above the Earth’s surface, for both the magnetic and velocity fields at the core surface, taking into account both magnetic diffusion and subgrid induction. We merge spatial information provided by numerical simulations, specifically from the Coupled Earth dynamo (CED) model (Aubert et al.(2013)Aubert, Finlay, & Fournier) and temporal constraints coming from a restriction of the field evolution to a chosen class of stochastic process. The sequential algorithm of BGA17, which considers as input data time series of spherical harmonic coefficients of the main field, is extended to account for both virtual observatory (Mandea & Olsen(2006)) and ground observatory time series that cover the period 1997–2017. Our approach has similarities with the previous works of Gillet et al.(2015a)Gillet, Barrois, & Finlay and Baerenzung et al.(2016)Baerenzung, Holschneider, & Lesur, which favoured flat flow spatial spectra at the CMB, since the spatial dynamo norm employed here departs from the norms often employed to ensure spectral convergence. In addition, our stochastic framework allows us to discuss posterior model errors for both the flow and the magnetic field.

The paper is organised as follows. In section §2 we describe the ground-based observatory data and satellite-based virtual observatory data, and the methodology we follow to recover magnetic and velocity fields at the CMB. In section §3.1, we present our resulting geomagnetic model and its associated uncertainties, before we analyse in §3.2 our core flow solutions. Finally, implications for our understanding of the core dynamics and possible further improvements for the algorithm are given in section §4.

2 Methodology

2.1 Ground-based and virtual observatory data

2.1.1 Ground observatory data

We use magnetic measurements made at 186 ground observatories (GO) covering the period 1997–2017. Hourly mean values are taken from the BGS database111ftp://ftp.nerc-murchison.ac.uk/geomag/Swarm/AUX_OBS, version 0111, using Intermagnet and WDC Edinburgh data as available in May 2017. The data have been checked and corrected for known baseline jumps (Macmillan & Olsen(2013)). ‘Revised monthly means’ were then derived from these hourly means, following the procedure described by Olsen et al.(2014)Olsen, Lühr, Finlay, Sabaka, Michaelis, Rauberg, & Tøffner-Clausen. Briefly, predictions of the large-scale magnetospheric field (and the associated induced field) from the CHAOS-6 field model, as well as predictions for the ionospheric Sq field (and the associated induced field) from the CM4 model (Sabaka et al.(2004)Sabaka, Olsen, & Purucker) are subtracted from the hourly mean values, and then robust (Huber-weighted) monthly mean values are computed using an iterative-reweighting procedure. Annual differences of such revised monthly means are routinely used in deriving the CHAOS series of field models and in order to study high resolution secular variation since, compared with simple monthly means, they are less contaminated by external field effects. Here, since we also wish to use the field itself for model construction, the median difference between each series and CHAOS-6 predictions was removed, in order to account in a simple way for the bias due to unmodelled crustal fields. In order to obtain the same sampling rate as that adopted for the virtual observatory series described below, the revised monthly mean series were finally averaged over 4 months windows to obtain the GO series used in our data assimilation scheme.

2.1.2 Virtual observatory data

In addition to GO data, we make use of satellite measurements from the CHAMP and Swarm missions covering respectively 2000–2010 and 2014–2017, through so-called virtual observatory (VO) data (Mandea & Olsen(2006); Olsen & Mandea(2007)). These provide a regular spatial and temporal sampling of the global field, convenient for our Kalman filter algorithm (detailed in §2.2) and involve estimates from an easily manageable number of locations, which has computational advantages.

VO data were computed using measurements collected by the CHAMP vector field magnetometer between July 2000 and September 2010 and from the Swarm vector field magnetometers, onboard all three satellites (Alpha, Bravo, Charlie), between January 2014 and April 2017. Starting from the CHAMP MAG-L3 and Swarm Level 1b MAG-L, version 0501, data products, we sub-sampled at 15s intervals the data in the vector field magnetometer (VFM) frame. Using the Euler rotation angles as given by the CHAOS-6-x3 model (which was based on Swarm and ground observation data up until April 2017222http://www.spacecenter.dk/files/magnetic-models/CHAOS-6/), we rotated the VFM data into an Earth-Centered Earth-Fixed (ECEF) coordinate frame.

Measurements from known problematic days were removed, for instance where satellite manoeuvres happened. Furthermore, gross data outliers with deviations more than 500 nT from CHAOS-6-x3 field model predictions were rejected. Based on previous studies of VO estimates (e.g., Beggan et al.(2009)Beggan, Whaler, & Macmillan), we then employed data selection criteria retaining only data for which:

  1. -

    the sun was at maximum 10superscript1010^{\circ}10 start_POSTSUPERSCRIPT ∘ end_POSTSUPERSCRIPT above horizon;

  2. -

    geomagnetic activity index Kp<3osubscriptKpsuperscript3o\mathrm{K_{p}}<3^{\mathrm{o}}roman_K start_POSTSUBSCRIPT roman_p end_POSTSUBSCRIPT < 3 start_POSTSUPERSCRIPT roman_o end_POSTSUPERSCRIPT;

  3. -

    the RC𝑅𝐶RCitalic_R italic_C disturbance index (Olsen et al.(2014)Olsen, Lühr, Finlay, Sabaka, Michaelis, Rauberg, & Tøffner-Clausen) had |dRC/dt|<3𝑑RC𝑑𝑡3|d\mathrm{RC}/dt|<3| italic_d roman_RC / italic_d italic_t | < 3 nT/h;

  4. -

    merging electric field at the magnetopause Em0.8subscriptEm0.8\mathrm{E_{m}}\leq 0.8roman_E start_POSTSUBSCRIPT roman_m end_POSTSUBSCRIPT ≤ 0.8 mV/m, with Em=0.33v4/3Bt2/3sin(|Θ|/2)subscript𝐸𝑚0.33superscript𝑣43superscriptsubscript𝐵𝑡23sinΘ2E_{m}=0.33v^{4/3}B_{t}^{2/3}\mathrm{sin}(|\Theta|/2)italic_E start_POSTSUBSCRIPT italic_m end_POSTSUBSCRIPT = 0.33 italic_v start_POSTSUPERSCRIPT 4 / 3 end_POSTSUPERSCRIPT italic_B start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 / 3 end_POSTSUPERSCRIPT roman_sin ( | roman_Θ | / 2 ). v𝑣vitalic_v is the solar wind speed, Θ=arctan(By/Bz)Θarctansubscript𝐵𝑦subscript𝐵𝑧\Theta=\mathrm{arctan}(B_{y}/B_{z})roman_Θ = roman_arctan ( italic_B start_POSTSUBSCRIPT italic_y end_POSTSUBSCRIPT / italic_B start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT ) and Bt=By2+Bz2subscript𝐵𝑡superscriptsubscript𝐵𝑦2superscriptsubscript𝐵𝑧2B_{t}=\sqrt{B_{y}^{2}+B_{z}^{2}}italic_B start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT = square-root start_ARG italic_B start_POSTSUBSCRIPT italic_y end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + italic_B start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG. Bysubscript𝐵𝑦B_{y}italic_B start_POSTSUBSCRIPT italic_y end_POSTSUBSCRIPT and Bzsubscript𝐵𝑧B_{z}italic_B start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT are components of the interplanetary magnetic field (IMF) in the geocentric solar magnetospheric (GSM) coordinate system, calculated using 2 hourly means of 1-min values of the IMF and solar wind extracted from the OMNI database333http://omniweb.gsfc.nasa.gov;

  5. -

    IMF Bz>0subscript𝐵𝑧0B_{z}>0italic_B start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT > 0 nT and IMF |By|<10subscript𝐵𝑦10|B_{y}|<10| italic_B start_POSTSUBSCRIPT italic_y end_POSTSUBSCRIPT | < 10 nT, again based on 2 hourly mean of 1 minute values.

Following this data selection, estimates of the fields due to various unmodelled sources were next removed from the data:

  1. (i)

    the magnetospheric and its induced fields as given by the CHAOS-6-x3 model;

  2. (ii)

    the ionospheric and its induced fields as given by the CM4 model (Sabaka et al.(2004)Sabaka, Olsen, & Purucker);

  3. (iii)

    the static internal field for spherical harmonic degrees n>20𝑛20n>20italic_n > 20 given by the CHAOS-6-x3 model.

Although imperfect, in our opinion it is more consistent to remove such estimates rather to ignore known field sources.

Based on this data we then carried robust inversions for time-averaged point estimates (i.e. VOs) using data windows of 4 months width (60 days each side of an epoch tjsubscript𝑡𝑗t_{j}italic_t start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT). In order to aid the robust inversion procedure in identifying and downweighting outliers, following Olsen & Mandea(2007) as a pre-processing step, we also removed a time-dependent internal field, here taken from the CHAOS-6-x3 model (Finlay et al.(2016b)Finlay, Olsen, Kotsiaros, Gillet, & Tøffner-Clausen), for spherical harmonic degrees 1 to 20, within each four month window. The CHAOS-6x-3 prediction at the target point and time was then added back at the end of the analysis. Note that this does not prevent our 4-monthly VO series, and the derived SV series from departing from CHAOS-6x-3; information about the time-dependence within each 4 month window is however lost.

We assume that the residual field 𝐁~~𝐁\tilde{\bf B}over~ start_ARG bold_B end_ARG, after the removal of the time-dependent internal field from the CHAOS-6-x3, can be represented as the gradient of a scalar potential V𝑉Vitalic_V, i.e. {linenomath*}

𝐁~=V.~𝐁𝑉\displaystyle\tilde{\bf B}=-\nabla V\,.over~ start_ARG bold_B end_ARG = - ∇ italic_V . (1)

The residual field and associated positions are transformed into a local Cartesian coordinate system with origin at the VO points of interest, with x𝑥xitalic_x pointing towards geographic South, y𝑦yitalic_y pointing towards East and z𝑧zitalic_z pointing upwards. We use an expansion of the local potential up to cubic terms. Because the geomagnetic field is irrotational (×𝐁~=0~𝐁0\nabla\times\tilde{\bf B}=0∇ × over~ start_ARG bold_B end_ARG = 0) and solenoidal (𝐁~=0~𝐁0\nabla\cdot\tilde{\bf B}=0∇ ⋅ over~ start_ARG bold_B end_ARG = 0), this local potential is entirely determined by 15 independent parameters: {linenomath*}

V(x,y,z)=vxx+vyy+vzz+vxxx2+vyyy2(vxx+vyy)z2𝑉𝑥𝑦𝑧subscript𝑣𝑥𝑥subscript𝑣𝑦𝑦subscript𝑣𝑧𝑧subscript𝑣𝑥𝑥superscript𝑥2subscript𝑣𝑦𝑦superscript𝑦2subscript𝑣𝑥𝑥subscript𝑣𝑦𝑦superscript𝑧2\displaystyle V(x,y,z)=v_{x}x+v_{y}y+v_{z}z+v_{xx}x^{2}+v_{yy}y^{2}-(v_{xx}+v_% {yy})z^{2}italic_V ( italic_x , italic_y , italic_z ) = italic_v start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT italic_x + italic_v start_POSTSUBSCRIPT italic_y end_POSTSUBSCRIPT italic_y + italic_v start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT italic_z + italic_v start_POSTSUBSCRIPT italic_x italic_x end_POSTSUBSCRIPT italic_x start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + italic_v start_POSTSUBSCRIPT italic_y italic_y end_POSTSUBSCRIPT italic_y start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT - ( italic_v start_POSTSUBSCRIPT italic_x italic_x end_POSTSUBSCRIPT + italic_v start_POSTSUBSCRIPT italic_y italic_y end_POSTSUBSCRIPT ) italic_z start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT (2)
+2vxyxy+2vxzxz+2vyzyz(vxyy+vxzz)x32subscript𝑣𝑥𝑦𝑥𝑦2subscript𝑣𝑥𝑧𝑥𝑧2subscript𝑣𝑦𝑧𝑦𝑧subscript𝑣𝑥𝑦𝑦subscript𝑣𝑥𝑧𝑧superscript𝑥3\displaystyle+2v_{xy}xy+2v_{xz}xz+2v_{yz}yz-(v_{xyy}+v_{xzz})x^{3}+ 2 italic_v start_POSTSUBSCRIPT italic_x italic_y end_POSTSUBSCRIPT italic_x italic_y + 2 italic_v start_POSTSUBSCRIPT italic_x italic_z end_POSTSUBSCRIPT italic_x italic_z + 2 italic_v start_POSTSUBSCRIPT italic_y italic_z end_POSTSUBSCRIPT italic_y italic_z - ( italic_v start_POSTSUBSCRIPT italic_x italic_y italic_y end_POSTSUBSCRIPT + italic_v start_POSTSUBSCRIPT italic_x italic_z italic_z end_POSTSUBSCRIPT ) italic_x start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT
+3vxxyx2y+3vxxzx2z+3vxyyxy2+3vxzzxz2+6vxyzxyz3subscript𝑣𝑥𝑥𝑦superscript𝑥2𝑦3subscript𝑣𝑥𝑥𝑧superscript𝑥2𝑧3subscript𝑣𝑥𝑦𝑦𝑥superscript𝑦23subscript𝑣𝑥𝑧𝑧𝑥superscript𝑧26subscript𝑣𝑥𝑦𝑧𝑥𝑦𝑧\displaystyle+3v_{xxy}x^{2}y+3v_{xxz}x^{2}z+3v_{xyy}xy^{2}+3v_{xzz}xz^{2}+6v_{% xyz}xyz+ 3 italic_v start_POSTSUBSCRIPT italic_x italic_x italic_y end_POSTSUBSCRIPT italic_x start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_y + 3 italic_v start_POSTSUBSCRIPT italic_x italic_x italic_z end_POSTSUBSCRIPT italic_x start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_z + 3 italic_v start_POSTSUBSCRIPT italic_x italic_y italic_y end_POSTSUBSCRIPT italic_x italic_y start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + 3 italic_v start_POSTSUBSCRIPT italic_x italic_z italic_z end_POSTSUBSCRIPT italic_x italic_z start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + 6 italic_v start_POSTSUBSCRIPT italic_x italic_y italic_z end_POSTSUBSCRIPT italic_x italic_y italic_z
(vxxyvyzz)y3+3vyyzy2z+3vyzzyz2(vxxz+vyyz)z3.subscript𝑣𝑥𝑥𝑦subscript𝑣𝑦𝑧𝑧superscript𝑦33subscript𝑣𝑦𝑦𝑧superscript𝑦2𝑧3subscript𝑣𝑦𝑧𝑧𝑦superscript𝑧2subscript𝑣𝑥𝑥𝑧subscript𝑣𝑦𝑦𝑧superscript𝑧3\displaystyle-(v_{xxy}-v_{yzz})y^{3}+3v_{yyz}y^{2}z+3v_{yzz}yz^{2}-(v_{xxz}+v_% {yyz})z^{3}\,.- ( italic_v start_POSTSUBSCRIPT italic_x italic_x italic_y end_POSTSUBSCRIPT - italic_v start_POSTSUBSCRIPT italic_y italic_z italic_z end_POSTSUBSCRIPT ) italic_y start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT + 3 italic_v start_POSTSUBSCRIPT italic_y italic_y italic_z end_POSTSUBSCRIPT italic_y start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_z + 3 italic_v start_POSTSUBSCRIPT italic_y italic_z italic_z end_POSTSUBSCRIPT italic_y italic_z start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT - ( italic_v start_POSTSUBSCRIPT italic_x italic_x italic_z end_POSTSUBSCRIPT + italic_v start_POSTSUBSCRIPT italic_y italic_y italic_z end_POSTSUBSCRIPT ) italic_z start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT .

For each VO position vector 𝐫k=(θk,ϕk,rk)subscript𝐫𝑘subscript𝜃𝑘subscriptitalic-ϕ𝑘subscript𝑟𝑘{\bf r}_{k}=(\theta_{k},\phi_{k},r_{k})bold_r start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT = ( italic_θ start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT , italic_ϕ start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT , italic_r start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT ) and at epoch tjsubscript𝑡𝑗t_{j}italic_t start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT, all data positioned within a cylinder of radius 850km (7.5absentsuperscript7.5\approx 7.5^{\circ}≈ 7.5 start_POSTSUPERSCRIPT ∘ end_POSTSUPERSCRIPT) of the VO target 𝐫ksubscript𝐫𝑘{\bf r}_{k}bold_r start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT, and within 60 days either side of tjsubscript𝑡𝑗t_{j}italic_t start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT were used to build a local data vector 𝐝k,jsuperscript𝐝𝑘𝑗{\bf d}^{k,j}bold_d start_POSTSUPERSCRIPT italic_k , italic_j end_POSTSUPERSCRIPT. These data are then related to the 15 parameters defining the VO potential model 𝐦vok,jsuperscriptsubscript𝐦𝑣𝑜𝑘𝑗{\bf m}_{vo}^{k,j}bold_m start_POSTSUBSCRIPT italic_v italic_o end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_k , italic_j end_POSTSUPERSCRIPT at that site and epoch via 𝐝k,j=𝐠¯¯k,j𝐦vok,jsuperscript𝐝𝑘𝑗superscript¯¯𝐠𝑘𝑗superscriptsubscript𝐦𝑣𝑜𝑘𝑗{\bf d}^{k,j}=\underline{\underline{{\bf g}}}^{k,j}{\bf m}_{vo}^{k,j}bold_d start_POSTSUPERSCRIPT italic_k , italic_j end_POSTSUPERSCRIPT = under¯ start_ARG under¯ start_ARG bold_g end_ARG end_ARG start_POSTSUPERSCRIPT italic_k , italic_j end_POSTSUPERSCRIPT bold_m start_POSTSUBSCRIPT italic_v italic_o end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_k , italic_j end_POSTSUPERSCRIPT, where the elements of the matrix 𝐠¯¯k,jsuperscript¯¯𝐠𝑘𝑗\underline{\underline{{\bf g}}}^{k,j}under¯ start_ARG under¯ start_ARG bold_g end_ARG end_ARG start_POSTSUPERSCRIPT italic_k , italic_j end_POSTSUPERSCRIPT are determined from (1) and (2).

Rather than working directly with 𝐝k,jsuperscript𝐝𝑘𝑗{\bf d}^{k,j}bold_d start_POSTSUPERSCRIPT italic_k , italic_j end_POSTSUPERSCRIPT in deriving 𝐦vok,jsuperscriptsubscript𝐦𝑣𝑜𝑘𝑗{\bf m}_{vo}^{k,j}bold_m start_POSTSUBSCRIPT italic_v italic_o end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_k , italic_j end_POSTSUPERSCRIPT we make use of along-track and East-West (using Swarm Alpha and Charlie only) sums and differences of the magnetic field components. An advantage of using field differences is that these have a reduced sensitivity to large-scale external signals, although data sums also need to be included in order to ensure sufficient information on the longer wavelengths core field. Using sums and differences has been found advantageous in a number of other field modelling efforts (Sabaka et al.(2015)Sabaka, Olsen, Tyler, & Kuvshinov; Olsen et al.(2015)Olsen, Hulot, Lesur, Finlay, Beggan, Chulliat, Sabaka, Floberghagen, Friis-Christensen, Haagmans, et al.). We calculate along-track (AT) sums (ΣΣ\Sigmaroman_Σ) and differences (ΔΔ\Deltaroman_Δ) as {linenomath*}

{ΣdiAT=[B~i(𝐫,t)+B~i(𝐫+δ𝐫,t+15s)]/2ΔdiAT=[B~i(𝐫,t)B~i(𝐫+δ𝐫,t+15s)].casesΣsuperscriptsubscript𝑑𝑖𝐴𝑇absentdelimited-[]subscript~𝐵𝑖𝐫𝑡subscript~𝐵𝑖𝐫𝛿𝐫𝑡15s2Δsuperscriptsubscript𝑑𝑖ATabsentdelimited-[]subscript~𝐵𝑖𝐫𝑡subscript~𝐵𝑖𝐫𝛿𝐫𝑡15s\displaystyle\left\{\begin{array}[]{rl}\Sigma d_{i}^{AT}=&[\tilde{B}_{i}(% \mathbf{r},t)+\tilde{B}_{i}(\mathbf{r}+\delta\mathbf{r},t+15\mathrm{s})]/2\\ \Delta d_{i}^{\mathrm{AT}}=&[\tilde{B}_{i}(\mathbf{r},t)-\tilde{B}_{i}(\mathbf% {r}+\delta\mathbf{r},t+15\mathrm{s})]\end{array}\right.\,.{ start_ARRAY start_ROW start_CELL roman_Σ italic_d start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_A italic_T end_POSTSUPERSCRIPT = end_CELL start_CELL [ over~ start_ARG italic_B end_ARG start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ( bold_r , italic_t ) + over~ start_ARG italic_B end_ARG start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ( bold_r + italic_δ bold_r , italic_t + 15 roman_s ) ] / 2 end_CELL end_ROW start_ROW start_CELL roman_Δ italic_d start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT roman_AT end_POSTSUPERSCRIPT = end_CELL start_CELL [ over~ start_ARG italic_B end_ARG start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ( bold_r , italic_t ) - over~ start_ARG italic_B end_ARG start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ( bold_r + italic_δ bold_r , italic_t + 15 roman_s ) ] end_CELL end_ROW end_ARRAY . (5)

B~i=𝟏i𝐁~(𝐫)subscript~𝐵𝑖subscript1𝑖~𝐁𝐫\tilde{B}_{i}={\mathbf{1}_{i}}\cdot\mathbf{\tilde{B}}(\mathbf{r})over~ start_ARG italic_B end_ARG start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT = bold_1 start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ⋅ over~ start_ARG bold_B end_ARG ( bold_r ) are the residual magnetic field components in spherical polar coordinates (where i=r,θ𝑖𝑟𝜃i=r,\thetaitalic_i = italic_r , italic_θ or ϕitalic-ϕ\phiitalic_ϕ, and 𝟏isubscript1𝑖{\mathbf{1}_{i}}bold_1 start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT are unit vectors). The East-West cross-track (CT) sums and differences between are calculated as {linenomath*}

{ΣdiCT=[B~iAlpha(𝐫1,t1)+B~iCharlie(𝐫2,t2)]/2ΔdiCT=[B~iAlpha(𝐫1,t1)B~iCharlie(𝐫2,t2)].casesΣsuperscriptsubscript𝑑𝑖CTabsentdelimited-[]superscriptsubscript~𝐵𝑖Alphasubscript𝐫1subscript𝑡1superscriptsubscript~𝐵𝑖Charliesubscript𝐫2subscript𝑡22Δsuperscriptsubscript𝑑𝑖CTabsentdelimited-[]superscriptsubscript~𝐵𝑖Alphasubscript𝐫1subscript𝑡1superscriptsubscript~𝐵𝑖Charliesubscript𝐫2subscript𝑡2\displaystyle\left\{\begin{array}[]{rl}\Sigma d_{i}^{\mathrm{CT}}=&[\tilde{B}_% {i}^{\mathrm{Alpha}}(\mathbf{r}_{1},t_{1})+\tilde{B}_{i}^{\mathrm{Charlie}}(% \mathbf{r}_{2},t_{2})]/2\\ \Delta d_{i}^{\mathrm{CT}}=&[\tilde{B}_{i}^{\mathrm{Alpha}}(\mathbf{r}_{1},t_{% 1})-\tilde{B}_{i}^{\mathrm{Charlie}}(\mathbf{r}_{2},t_{2})]\end{array}\right.\,.{ start_ARRAY start_ROW start_CELL roman_Σ italic_d start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT roman_CT end_POSTSUPERSCRIPT = end_CELL start_CELL [ over~ start_ARG italic_B end_ARG start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT roman_Alpha end_POSTSUPERSCRIPT ( bold_r start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT , italic_t start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ) + over~ start_ARG italic_B end_ARG start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT roman_Charlie end_POSTSUPERSCRIPT ( bold_r start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT , italic_t start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ) ] / 2 end_CELL end_ROW start_ROW start_CELL roman_Δ italic_d start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT roman_CT end_POSTSUPERSCRIPT = end_CELL start_CELL [ over~ start_ARG italic_B end_ARG start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT roman_Alpha end_POSTSUPERSCRIPT ( bold_r start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT , italic_t start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ) - over~ start_ARG italic_B end_ARG start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT roman_Charlie end_POSTSUPERSCRIPT ( bold_r start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT , italic_t start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ) ] end_CELL end_ROW end_ARRAY . (8)

Here, for a given orbit of Alpha we select the corresponding Charlie data to be the one closest in colatitude such that |δt|=|t1t2|<50s𝛿𝑡subscript𝑡1subscript𝑡250𝑠|\delta t|=|t_{1}-t_{2}|<50s| italic_δ italic_t | = | italic_t start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT - italic_t start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT | < 50 italic_s. Crucially, in order to relate these sums and differences to the VO model parameters, we also take sums and differences of the elements of the design matrices 𝐠¯¯k,jsuperscript¯¯𝐠𝑘𝑗\underline{\underline{{\bf g}}}^{k,j}under¯ start_ARG under¯ start_ARG bold_g end_ARG end_ARG start_POSTSUPERSCRIPT italic_k , italic_j end_POSTSUPERSCRIPT associated with the predictions of the VO model for the field components at the individual data locations. This results in a design matrix {linenomath*}

𝐆¯¯k,j=[Σ𝐠¯¯k,jΔ𝐠¯¯k,j]superscript¯¯𝐆𝑘𝑗matrixsuperscript¯¯Σ𝐠𝑘𝑗superscript¯¯Δ𝐠𝑘𝑗\underline{\underline{{\bf G}}}^{k,j}=\begin{bmatrix}\underline{\underline{% \Sigma{\bf g}}}^{k,j}\\ \underline{\underline{\Delta{\bf g}}}^{k,j}\end{bmatrix}under¯ start_ARG under¯ start_ARG bold_G end_ARG end_ARG start_POSTSUPERSCRIPT italic_k , italic_j end_POSTSUPERSCRIPT = [ start_ARG start_ROW start_CELL under¯ start_ARG under¯ start_ARG roman_Σ bold_g end_ARG end_ARG start_POSTSUPERSCRIPT italic_k , italic_j end_POSTSUPERSCRIPT end_CELL end_ROW start_ROW start_CELL under¯ start_ARG under¯ start_ARG roman_Δ bold_g end_ARG end_ARG start_POSTSUPERSCRIPT italic_k , italic_j end_POSTSUPERSCRIPT end_CELL end_ROW end_ARG ] (9)

associated with the data vector 𝐃k,j=[Σ𝐝k,jΔ𝐝k,j]Tsuperscript𝐃𝑘𝑗superscriptdelimited-[]Σsuperscript𝐝𝑘𝑗Δsuperscript𝐝𝑘𝑗𝑇{\bf D}^{k,j}=\left[\Sigma{\bf d}^{k,j}\,\,\Delta{\bf d}^{k,j}\right]^{T}bold_D start_POSTSUPERSCRIPT italic_k , italic_j end_POSTSUPERSCRIPT = [ roman_Σ bold_d start_POSTSUPERSCRIPT italic_k , italic_j end_POSTSUPERSCRIPT roman_Δ bold_d start_POSTSUPERSCRIPT italic_k , italic_j end_POSTSUPERSCRIPT ] start_POSTSUPERSCRIPT italic_T end_POSTSUPERSCRIPT. In this way we fully account for the change in the unit vectors associated with the two locations contributing to the sums and differences when deriving the parameters 𝐦vok,jsuperscriptsubscript𝐦𝑣𝑜𝑘𝑗{\bf m}_{vo}^{k,j}bold_m start_POSTSUBSCRIPT italic_v italic_o end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_k , italic_j end_POSTSUPERSCRIPT. The inversion for each 𝐦vok,jsuperscriptsubscript𝐦𝑣𝑜𝑘𝑗{\bf m}_{vo}^{k,j}bold_m start_POSTSUBSCRIPT italic_v italic_o end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_k , italic_j end_POSTSUPERSCRIPT is carried out via a robust Huber weighted least-square fit {linenomath*}

𝐦vok,j=[(𝐆¯¯k,j)T𝐖𝐆¯¯k,j]1(𝐆¯¯k,j)T𝐃k,jsuperscriptsubscript𝐦𝑣𝑜𝑘𝑗superscriptdelimited-[]superscriptsuperscript¯¯𝐆𝑘𝑗𝑇𝐖superscript¯¯𝐆𝑘𝑗1superscriptsuperscript¯¯𝐆𝑘𝑗𝑇superscript𝐃𝑘𝑗\displaystyle{\bf m}_{vo}^{k,j}=\left[(\underline{\underline{{\bf G}}}^{k,j})^% {T}{\bf W}\underline{\underline{{\bf G}}}^{k,j}\right]^{-1}(\underline{% \underline{{\bf G}}}^{k,j})^{T}{\bf D}^{k,j}bold_m start_POSTSUBSCRIPT italic_v italic_o end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_k , italic_j end_POSTSUPERSCRIPT = [ ( under¯ start_ARG under¯ start_ARG bold_G end_ARG end_ARG start_POSTSUPERSCRIPT italic_k , italic_j end_POSTSUPERSCRIPT ) start_POSTSUPERSCRIPT italic_T end_POSTSUPERSCRIPT bold_W under¯ start_ARG under¯ start_ARG bold_G end_ARG end_ARG start_POSTSUPERSCRIPT italic_k , italic_j end_POSTSUPERSCRIPT ] start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT ( under¯ start_ARG under¯ start_ARG bold_G end_ARG end_ARG start_POSTSUPERSCRIPT italic_k , italic_j end_POSTSUPERSCRIPT ) start_POSTSUPERSCRIPT italic_T end_POSTSUPERSCRIPT bold_D start_POSTSUPERSCRIPT italic_k , italic_j end_POSTSUPERSCRIPT (10)

where 𝐖𝐖{\bf W}bold_W is a diagonal vector of Huber weights that ensures a robust solution (Olsen(2002); Sabaka et al.(2004)Sabaka, Olsen, & Purucker) and are iteratively updated until convergence. Once 𝐦vok,jsuperscriptsubscript𝐦𝑣𝑜𝑘𝑗{\bf m}_{vo}^{k,j}bold_m start_POSTSUBSCRIPT italic_v italic_o end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_k , italic_j end_POSTSUPERSCRIPT is determined, the three field components at the site and epoch of interest, 𝐁~k(𝐫k,tj)=Vk(𝐫k,tj)subscript~𝐁𝑘subscript𝐫𝑘subscript𝑡𝑗subscript𝑉𝑘subscript𝐫𝑘subscript𝑡𝑗\tilde{\bf B}_{k}({\bf r}_{k},t_{j})=-\nabla V_{k}({\bf r}_{k},t_{j})over~ start_ARG bold_B end_ARG start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT ( bold_r start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT , italic_t start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ) = - ∇ italic_V start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT ( bold_r start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT , italic_t start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ), are computed and added back on to the CHAOS-6-x3 prediction for the internal field (for degrees 1-14 only, to avoid as far as possible the lithospheric field) at the target location.

We constructed VO estimates at PVO=200subscript𝑃VO200P_{\mathrm{VO}}=200italic_P start_POSTSUBSCRIPT roman_VO end_POSTSUBSCRIPT = 200 locations, with a spacing of about 1600 km (14absentsuperscript14\approx 14^{\circ}≈ 14 start_POSTSUPERSCRIPT ∘ end_POSTSUPERSCRIPT, see dots in Figure 1), located in an approximately equal area grid based on the spherical surface partition algorithm of Leopardi(2006). The altitude of the VOs are 300km and 500km during the CHAMP and Swarm periods, respectively. Using predictions of the three components (Brsubscript𝐵𝑟B_{r}italic_B start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT, Bθsubscript𝐵𝜃B_{\theta}italic_B start_POSTSUBSCRIPT italic_θ end_POSTSUBSCRIPT, Bϕsubscript𝐵italic-ϕB_{\phi}italic_B start_POSTSUBSCRIPT italic_ϕ end_POSTSUBSCRIPT) of the magnetic field at PVOsubscript𝑃VOP_{\mathrm{VO}}italic_P start_POSTSUBSCRIPT roman_VO end_POSTSUBSCRIPT locations, we finally obtain 3PVO3subscript𝑃VO3P_{\mathrm{VO}}3 italic_P start_POSTSUBSCRIPT roman_VO end_POSTSUBSCRIPT time series (i.e. one point every 4 months during CHAMP and Swarm times, 48 epochs in all), stored in a vector 𝐲VO(t)subscript𝐲VO𝑡{\bf y}_{\mathrm{VO}}(t)bold_y start_POSTSUBSCRIPT roman_VO end_POSTSUBSCRIPT ( italic_t ). The SV was computed as annual differences of the 4 month time series.

2.1.3 Uncertainty estimates for the GO and VO series

In order to obtain as much information as possible from the GO and VO data, while at the same time seeking to avoid overfitting them, it is important that appropriate uncertainty estimates are specified for each time series. We define 𝐂GOsubscript𝐂GO{\bf C}_{\mathrm{GO}}bold_C start_POSTSUBSCRIPT roman_GO end_POSTSUBSCRIPT and 𝐂VOsubscript𝐂VO{\bf C}_{\mathrm{VO}}bold_C start_POSTSUBSCRIPT roman_VO end_POSTSUBSCRIPT to be the measurement error cross-covariance matrices for GO and VO data at each epoch, of sizes respectively 3PGO×3PGO3subscript𝑃𝐺𝑂3subscript𝑃𝐺𝑂3P_{GO}\times 3P_{GO}3 italic_P start_POSTSUBSCRIPT italic_G italic_O end_POSTSUBSCRIPT × 3 italic_P start_POSTSUBSCRIPT italic_G italic_O end_POSTSUBSCRIPT and 3PVO×3PVO3subscript𝑃𝑉𝑂3subscript𝑃𝑉𝑂3P_{VO}\times 3P_{VO}3 italic_P start_POSTSUBSCRIPT italic_V italic_O end_POSTSUBSCRIPT × 3 italic_P start_POSTSUBSCRIPT italic_V italic_O end_POSTSUBSCRIPT. Data errors are supposed independent of time. Different data uncertainties are assigned for the VO’s derived from CHAMP and Swarm respectively.

Regarding the GO time series described above, we follow a similar approach to that used in CHAOS field model series (Olsen et al.(2014)Olsen, Lühr, Finlay, Sabaka, Michaelis, Rauberg, & Tøffner-Clausen; Finlay et al.(2016b)Finlay, Olsen, Kotsiaros, Gillet, & Tøffner-Clausen) and derive uncertainty estimates as follows. A three-by three covariance matrix was computed for each observatory location from the time-series of the three components, after removing the predictions of the CHAOS-6 field model and de-trending. The square root of the diagonal elements of these covariance matrices were taken to be the uncertainty estimates for each component at each observatory. The same procedure was applied to both the MF and SV series.

For consistency, a very similar procedure was also applied to the VO series in order to obtain their uncertainty estimates. For each VO location, covariances were calculated between the time series of the three components (after removing from each series the predictions of the CHAOS-6 model and de-trending), in order to obtain a three-by three covariance matrix. A robust procedure for calculating the covariances (using the Minimum Covariance Determinant estimator, Verboven & Hubert(2005)) was employed. However, only the square-root of the diagonal elements of the covariance matrices were taken to be the uncertainty estimates for each series, with similar procedures applied to both MF and SV series. To illustrate the range of the adopted uncertainty estimates, we show in Figure 1 the r.m.s. SV uncertainty estimates for all locations where data (GO or VO) are used in this study.

Refer to caption

Refer to caption

Refer to caption

Figure 1: SV observation error estimates (colorscale in nT/y𝑛𝑇𝑦nT/yitalic_n italic_T / italic_y) at all location where GOs (hexagons) and VOs (circles) are used in this study, for the 3 components B˙rsubscript˙𝐵𝑟\dot{B}_{r}over˙ start_ARG italic_B end_ARG start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT, B˙θsubscript˙𝐵𝜃\dot{B}_{\theta}over˙ start_ARG italic_B end_ARG start_POSTSUBSCRIPT italic_θ end_POSTSUBSCRIPT and B˙ϕsubscript˙𝐵italic-ϕ\dot{B}_{\phi}over˙ start_ARG italic_B end_ARG start_POSTSUBSCRIPT italic_ϕ end_POSTSUBSCRIPT (from top to bottom). The size of the markers is proportional to the magnitude of the a-priori error estimates.

Note that by using only the diagonal elements of 𝐂GOsubscript𝐂GO{\bf C}_{\mathrm{GO}}bold_C start_POSTSUBSCRIPT roman_GO end_POSTSUBSCRIPT and 𝐂VOsubscript𝐂VO{\bf C}_{\mathrm{VO}}bold_C start_POSTSUBSCRIPT roman_VO end_POSTSUBSCRIPT we effectively consider the errors on each GO and VO series to be uncorrelated with the errors on other series. In reality errors between components and between series will be correlated. This can be taken into account using full (i.e. dense) covariance matrices. It is however challenging to estimate cross-covariances for matrices of size larger than the length of the contributing times series (consisting of one sample every four months). We therefore postpone this step to future studies. Instead, by restricting to only 200 VO locations and ensuring that there was little overlap between the VO search radii we reduce as far as possible the correlations between distinct VO series.

Finally, we concatenate the above GO and VO main field data vectors for each epoch into 𝐲o(t)=[𝐲VOT𝐲GOT]Tsuperscript𝐲𝑜𝑡superscriptdelimited-[]superscriptsubscript𝐲VO𝑇superscriptsubscript𝐲GO𝑇𝑇{\bf y}^{o}(t)=[{\bf y}_{\mathrm{VO}}^{T}\,{\bf y}_{\mathrm{GO}}^{T}]^{T}bold_y start_POSTSUPERSCRIPT italic_o end_POSTSUPERSCRIPT ( italic_t ) = [ bold_y start_POSTSUBSCRIPT roman_VO end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_T end_POSTSUPERSCRIPT bold_y start_POSTSUBSCRIPT roman_GO end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_T end_POSTSUPERSCRIPT ] start_POSTSUPERSCRIPT italic_T end_POSTSUPERSCRIPT. The associated observation errors covariance matrix 𝐑yysubscript𝐑𝑦𝑦{\bf R}_{yy}bold_R start_POSTSUBSCRIPT italic_y italic_y end_POSTSUBSCRIPT, of rank P=3PVO+3PGO𝑃3subscript𝑃VO3subscript𝑃GOP=3P_{\mathrm{VO}}+3P_{\mathrm{GO}}italic_P = 3 italic_P start_POSTSUBSCRIPT roman_VO end_POSTSUBSCRIPT + 3 italic_P start_POSTSUBSCRIPT roman_GO end_POSTSUBSCRIPT, is thus derived from the diagonals of 𝐂VOsubscript𝐂VO{\bf C}_{\mathrm{VO}}bold_C start_POSTSUBSCRIPT roman_VO end_POSTSUBSCRIPT and 𝐂GOsubscript𝐂GO{\bf C}_{\mathrm{GO}}bold_C start_POSTSUBSCRIPT roman_GO end_POSTSUBSCRIPT. In the next section we will consider both main field and secular variation data. SV data 𝐲˙o(t)superscript˙𝐲𝑜𝑡\dot{\bf y}^{o}(t)over˙ start_ARG bold_y end_ARG start_POSTSUPERSCRIPT italic_o end_POSTSUPERSCRIPT ( italic_t ) are computed as annual differences of the four monthly (GO or VO) series. We follow the same approach as above to estimate the SV data errors variances (shown in Figure 1) that are stored in a diagonal matrix 𝐑y˙y˙subscript𝐑˙𝑦˙𝑦{\bf R}_{\dot{y}\dot{y}}bold_R start_POSTSUBSCRIPT over˙ start_ARG italic_y end_ARG over˙ start_ARG italic_y end_ARG end_POSTSUBSCRIPT of rank P𝑃Pitalic_P.

2.2 Re-analysis of GO and VO data ground and satellite magnetic observations

The assimilation algorithm used in the present study is essentially the one derived by BGA17 (see their table 2 for a summary). It is a sequential tool, consisting of a succession of forecast and analysis steps. The main modifications concern the direct integration of observations at and above the Earth’s surface, while BGA17 considered data in the form of MF and SV spherical harmonic coefficients. We begin by recalling the main points of our stochastic forecast model, before we go on to describe the changes implemented in the present study regarding the analysis step. These essentially concern the observation operator linking the state variables to the observations.

2.2.1 Stochastic forecast model

We forecast the evolution of the radial magnetic field, Brsubscript𝐵𝑟B_{r}italic_B start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT, at the CMB using the radial component of the induction equation, written as {linenomath*}

B¯rt=h(𝒖HB¯r)¯+er+dr(𝒖H,B¯r),subscript¯𝐵𝑟𝑡¯subscriptsubscript𝒖𝐻subscript¯𝐵𝑟subscript𝑒𝑟subscript𝑑𝑟subscript𝒖𝐻subscript¯𝐵𝑟\displaystyle\frac{\partial\overline{B}_{r}}{\partial t}=-\overline{\nabla_{h}% \cdot\left({\bm{u}}_{H}\overline{B}_{r}\right)}+e_{r}+d_{r}({\bm{u}}_{H},% \overline{B}_{r})\,,divide start_ARG ∂ over¯ start_ARG italic_B end_ARG start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT end_ARG start_ARG ∂ italic_t end_ARG = - over¯ start_ARG ∇ start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT ⋅ ( bold_italic_u start_POSTSUBSCRIPT italic_H end_POSTSUBSCRIPT over¯ start_ARG italic_B end_ARG start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT ) end_ARG + italic_e start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT + italic_d start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT ( bold_italic_u start_POSTSUBSCRIPT italic_H end_POSTSUBSCRIPT , over¯ start_ARG italic_B end_ARG start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT ) , (11)

where overlines mean the projection onto large length-scales. ersubscript𝑒𝑟e_{r}italic_e start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT stands for the subgrid induction processes arising due to the unresolved magnetic field at small length-scales, 𝒖Hsubscript𝒖𝐻{\bm{u}}_{H}bold_italic_u start_POSTSUBSCRIPT italic_H end_POSTSUBSCRIPT is the horizontal flow, and drsubscript𝑑𝑟d_{r}italic_d start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT, enslaved to B¯rsubscript¯𝐵𝑟\overline{B}_{r}over¯ start_ARG italic_B end_ARG start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT and 𝒖Hsubscript𝒖𝐻{\bm{u}}_{H}bold_italic_u start_POSTSUBSCRIPT italic_H end_POSTSUBSCRIPT, approximates the radial component of the diffusion operator (see below). The evolutions of ersubscript𝑒𝑟e_{r}italic_e start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT and 𝒖Hsubscript𝒖𝐻{\bm{u}}_{H}bold_italic_u start_POSTSUBSCRIPT italic_H end_POSTSUBSCRIPT are governed by order one auto-regressive stochastic processes, {linenomath*}

derdt+erτe=ζe,dsubscript𝑒𝑟d𝑡subscript𝑒𝑟subscript𝜏𝑒subscript𝜁𝑒\displaystyle\frac{\mathrm{d}e_{r}}{\mathrm{d}t}+\frac{e_{r}}{\tau_{e}}=\zeta_% {e}\,,divide start_ARG roman_d italic_e start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT end_ARG start_ARG roman_d italic_t end_ARG + divide start_ARG italic_e start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT end_ARG start_ARG italic_τ start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT end_ARG = italic_ζ start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT , (12)
{linenomath*}
d𝒖Hdt+(𝒖H𝒖^H)τu=ζu,dsubscript𝒖𝐻d𝑡subscript𝒖𝐻subscript^𝒖𝐻subscript𝜏𝑢subscript𝜁𝑢\displaystyle\frac{\mathrm{d}{\bm{u}}_{H}}{\mathrm{d}t}+\frac{({\bm{u}}_{H}-% \hat{\bm{u}}_{H})}{\tau_{u}}=\zeta_{u}\,,divide start_ARG roman_d bold_italic_u start_POSTSUBSCRIPT italic_H end_POSTSUBSCRIPT end_ARG start_ARG roman_d italic_t end_ARG + divide start_ARG ( bold_italic_u start_POSTSUBSCRIPT italic_H end_POSTSUBSCRIPT - over^ start_ARG bold_italic_u end_ARG start_POSTSUBSCRIPT italic_H end_POSTSUBSCRIPT ) end_ARG start_ARG italic_τ start_POSTSUBSCRIPT italic_u end_POSTSUBSCRIPT end_ARG = italic_ζ start_POSTSUBSCRIPT italic_u end_POSTSUBSCRIPT , (13)

with ζesubscript𝜁𝑒\zeta_{e}italic_ζ start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT and ζusubscript𝜁𝑢\zeta_{u}italic_ζ start_POSTSUBSCRIPT italic_u end_POSTSUBSCRIPT white noise processes, and 𝒖^Hsubscript^𝒖𝐻\hat{\bm{u}}_{H}over^ start_ARG bold_italic_u end_ARG start_POSTSUBSCRIPT italic_H end_POSTSUBSCRIPT the background flow model (obtained as the time-averaged flow from the CED model). These processes come from the same family of process as employed by Baerenzung et al.(2017)Baerenzung, Holschneider, Wicht, Sanchez, & Lesur. For each process, an effective restoring force is implemented via single time scales that we respectively fix as τe=10subscript𝜏𝑒10\tau_{e}=10italic_τ start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT = 10 yrs and τu=30subscript𝜏𝑢30\tau_{u}=30italic_τ start_POSTSUBSCRIPT italic_u end_POSTSUBSCRIPT = 30 yrs. Spatial cross-covariances of the two above fields are derived from statistics of a free run of the CED (Aubert et al.(2013)Aubert, Finlay, & Fournier).

The advected fields ersubscript𝑒𝑟e_{r}italic_e start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT, 𝒖Hsubscript𝒖𝐻{\bm{u}}_{H}bold_italic_u start_POSTSUBSCRIPT italic_H end_POSTSUBSCRIPT, B¯rsubscript¯𝐵𝑟\overline{B}_{r}over¯ start_ARG italic_B end_ARG start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT and drsubscript𝑑𝑟d_{r}italic_d start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT are represented through spherical harmonics, whose coefficients are stored in vectors 𝐞(t)𝐞𝑡{\bf e}(t)bold_e ( italic_t ), 𝐮(t)𝐮𝑡{\bf u}(t)bold_u ( italic_t ), 𝐛(t)𝐛𝑡{\bf b}(t)bold_b ( italic_t ) and 𝐝(t)𝐝𝑡{\bf d}(t)bold_d ( italic_t ), respectively. Diffusion in equation (11), and its dependence on ersubscript𝑒𝑟e_{r}italic_e start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT and 𝐮Hsubscript𝐮𝐻{\bf u}_{H}bold_u start_POSTSUBSCRIPT italic_H end_POSTSUBSCRIPT, is also an expression of cross-covariances extracted from the CED (involving the radial magnetic field below the CMB). The projection onto large length-scales is processed in the spectral domain, restricting the induction equation (and thus the expansion of the fields ersubscript𝑒𝑟e_{r}italic_e start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT, B¯rsubscript¯𝐵𝑟\overline{B}_{r}over¯ start_ARG italic_B end_ARG start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT and drsubscript𝑑𝑟d_{r}italic_d start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT) to spherical harmonic degrees nnb=14𝑛subscript𝑛𝑏14n\leq n_{b}=14italic_n ≤ italic_n start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT = 14, while the velocity field is truncated at nu=18subscript𝑛𝑢18n_{u}=18italic_n start_POSTSUBSCRIPT italic_u end_POSTSUBSCRIPT = 18. We write as 𝐛˙(t)˙𝐛𝑡\dot{\bf b}(t)over˙ start_ARG bold_b end_ARG ( italic_t ) the vector of SV spherical harmonic coefficients.

2.2.2 Integrating ground and satellite data in the assimilation tool

We write as 𝐌𝐌{\bf M}bold_M the operator that links the vector 𝐛(t)𝐛𝑡{\bf b}(t)bold_b ( italic_t ) to the three components main field observations 𝐲(t)𝐲𝑡{\bf y}(t)bold_y ( italic_t ) in the spatial domain (e.g. Olsen et al.(2010)Olsen, Glassmeier, & Jia): {linenomath*}

𝐲(t)=𝐌𝐛(t).𝐲𝑡𝐌𝐛𝑡\displaystyle{\bf y}(t)={\bf M}{\bf b}(t)\,.bold_y ( italic_t ) = bold_Mb ( italic_t ) . (14)

At each epoch it is of size no×nb(nb+2)subscript𝑛𝑜subscript𝑛𝑏subscript𝑛𝑏2n_{o}\times n_{b}(n_{b}+2)italic_n start_POSTSUBSCRIPT italic_o end_POSTSUBSCRIPT × italic_n start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT ( italic_n start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT + 2 ), with no=3(PVO+PGO)subscript𝑛𝑜3subscript𝑃VOsubscript𝑃GOn_{o}=3(P_{\mathrm{VO}}+P_{\mathrm{GO}})italic_n start_POSTSUBSCRIPT italic_o end_POSTSUBSCRIPT = 3 ( italic_P start_POSTSUBSCRIPT roman_VO end_POSTSUBSCRIPT + italic_P start_POSTSUBSCRIPT roman_GO end_POSTSUBSCRIPT ) the size of the observation vector. The matrix 𝐌𝐌{\bf M}bold_M is composed of sub-matrices 𝐌rsubscript𝐌𝑟{\bf M}_{r}bold_M start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT, 𝐌θsubscript𝐌𝜃{\bf M}_{\theta}bold_M start_POSTSUBSCRIPT italic_θ end_POSTSUBSCRIPT and 𝐌ϕsubscript𝐌italic-ϕ{\bf M}_{\phi}bold_M start_POSTSUBSCRIPT italic_ϕ end_POSTSUBSCRIPT, depending on the considered component of the magnetic field. In practice, elements of the matrix are, for a column j𝑗jitalic_j corresponding to a coefficient gnjmjsuperscriptsubscript𝑔subscript𝑛𝑗subscript𝑚𝑗g_{n_{j}}^{m_{j}}italic_g start_POSTSUBSCRIPT italic_n start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_m start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT end_POSTSUPERSCRIPT, and a line i𝑖iitalic_i to an observation at a coordinate 𝐫i=(ri,θi,ϕi)subscript𝐫𝑖subscript𝑟𝑖subscript𝜃𝑖subscriptitalic-ϕ𝑖{\bf r}_{i}=(r_{i},\theta_{i},\phi_{i})bold_r start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT = ( italic_r start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT , italic_θ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT , italic_ϕ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ), {linenomath*}

𝐌ri,jsubscript𝐌𝑟𝑖𝑗\displaystyle{\bf M}_{ri,j}bold_M start_POSTSUBSCRIPT italic_r italic_i , italic_j end_POSTSUBSCRIPT =\displaystyle== (nj+1)(ari)nj+2𝒫nm(θi)cos(mjϕi),subscript𝑛𝑗1superscriptsuperscript𝑎direct-sumsubscript𝑟𝑖subscript𝑛𝑗2superscriptsubscript𝒫𝑛𝑚subscript𝜃𝑖subscript𝑚𝑗subscriptitalic-ϕ𝑖\displaystyle(n_{j}+1)\left(\dfrac{a^{\oplus}}{r_{i}}\right)^{n_{j}+2}\mathscr% {P}_{n}^{m}(\theta_{i})\cos(m_{j}\phi_{i})\,,( italic_n start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT + 1 ) ( divide start_ARG italic_a start_POSTSUPERSCRIPT ⊕ end_POSTSUPERSCRIPT end_ARG start_ARG italic_r start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT end_ARG ) start_POSTSUPERSCRIPT italic_n start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT + 2 end_POSTSUPERSCRIPT script_P start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_m end_POSTSUPERSCRIPT ( italic_θ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ) roman_cos ( italic_m start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT italic_ϕ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ) , (15)
𝐌θi,jsubscript𝐌𝜃𝑖𝑗\displaystyle{\bf M}_{\theta i,j}bold_M start_POSTSUBSCRIPT italic_θ italic_i , italic_j end_POSTSUBSCRIPT =\displaystyle== (ari)nj+2d𝒫nm(θi)dθcos(mjϕi),superscriptsuperscript𝑎direct-sumsubscript𝑟𝑖subscript𝑛𝑗2𝑑superscriptsubscript𝒫𝑛𝑚subscript𝜃𝑖𝑑𝜃subscript𝑚𝑗subscriptitalic-ϕ𝑖\displaystyle\left(\dfrac{a^{\oplus}}{r_{i}}\right)^{n_{j}+2}\dfrac{d\mathscr{% P}_{n}^{m}(\theta_{i})}{d\theta}\cos(m_{j}\phi_{i})\,,( divide start_ARG italic_a start_POSTSUPERSCRIPT ⊕ end_POSTSUPERSCRIPT end_ARG start_ARG italic_r start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT end_ARG ) start_POSTSUPERSCRIPT italic_n start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT + 2 end_POSTSUPERSCRIPT divide start_ARG italic_d script_P start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_m end_POSTSUPERSCRIPT ( italic_θ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ) end_ARG start_ARG italic_d italic_θ end_ARG roman_cos ( italic_m start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT italic_ϕ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ) , (16)
𝐌ϕi,jsubscript𝐌italic-ϕ𝑖𝑗\displaystyle{\bf M}_{\phi i,j}bold_M start_POSTSUBSCRIPT italic_ϕ italic_i , italic_j end_POSTSUBSCRIPT =\displaystyle== (ari)nj+2mj𝒫nm(θi)sin(θi)(1)sin(mjϕi).superscriptsuperscript𝑎direct-sumsubscript𝑟𝑖subscript𝑛𝑗2subscript𝑚𝑗superscriptsubscript𝒫𝑛𝑚subscript𝜃𝑖subscript𝜃𝑖1subscript𝑚𝑗subscriptitalic-ϕ𝑖\displaystyle\left(\dfrac{a^{\oplus}}{r_{i}}\right)^{n_{j}+2}\dfrac{m_{j}% \mathscr{P}_{n}^{m}(\theta_{i})}{\sin(\theta_{i})}(-1)\sin(m_{j}\phi_{i})\,.( divide start_ARG italic_a start_POSTSUPERSCRIPT ⊕ end_POSTSUPERSCRIPT end_ARG start_ARG italic_r start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT end_ARG ) start_POSTSUPERSCRIPT italic_n start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT + 2 end_POSTSUPERSCRIPT divide start_ARG italic_m start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT script_P start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_m end_POSTSUPERSCRIPT ( italic_θ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ) end_ARG start_ARG roman_sin ( italic_θ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ) end_ARG ( - 1 ) roman_sin ( italic_m start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT italic_ϕ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ) . (17)

For a line j𝑗jitalic_j corresponding to a coefficient hnjmjsuperscriptsubscriptsubscript𝑛𝑗subscript𝑚𝑗h_{n_{j}}^{m_{j}}italic_h start_POSTSUBSCRIPT italic_n start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_m start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT end_POSTSUPERSCRIPT, the function sin\sinroman_sin replaces cos\cosroman_cos in (15) and (16), and cos\cosroman_cos replaces (1)sin1(-1)\sin( - 1 ) roman_sin in (17). a=6371.2superscript𝑎direct-sum6371.2a^{\oplus}=6371.2italic_a start_POSTSUPERSCRIPT ⊕ end_POSTSUPERSCRIPT = 6371.2 km is the Earth’s spherical reference radius and 𝒫nmsuperscriptsubscript𝒫𝑛𝑚\mathscr{P}_{n}^{m}script_P start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_m end_POSTSUPERSCRIPT are the Legendre polynomials.

The analysis in the Kalman filter algorithm employed by BGA17 consists of two steps: first an analysis of the vector 𝐛𝐛{\bf b}bold_b containing MF spherical harmonic coefficients from MF spherical harmonic coefficients data, and second an analysis of the vector 𝐳𝐳{\bf z}bold_z (that concatenates 𝐮𝐮{\bf u}bold_u and 𝐞𝐞{\bf e}bold_e) from SV spherical harmonic coefficients data. Writing as 𝐏bbfsuperscriptsubscript𝐏𝑏𝑏𝑓{\bf P}_{bb}^{f}bold_P start_POSTSUBSCRIPT italic_b italic_b end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_f end_POSTSUPERSCRIPT the forecast model covariance matrix for 𝐛𝐛{\bf b}bold_b, the first analysis (equation (19) of BGA17) is replaced here by {linenomath*}

k[1,Nm],for-all𝑘1subscript𝑁𝑚\displaystyle\forall k\in[1,N_{m}],\;∀ italic_k ∈ [ 1 , italic_N start_POSTSUBSCRIPT italic_m end_POSTSUBSCRIPT ] , (18)
𝐛ka(ta)=𝐛kf(ta)+𝐏bbf𝐌T[𝐌𝐏bbf𝐌T+𝐑yy]1superscript𝐛𝑘𝑎subscript𝑡𝑎superscript𝐛𝑘𝑓subscript𝑡𝑎superscriptsubscript𝐏𝑏𝑏𝑓superscript𝐌𝑇superscriptdelimited-[]superscriptsubscript𝐌𝐏𝑏𝑏𝑓superscript𝐌𝑇subscript𝐑𝑦𝑦1\displaystyle{\bf b}^{ka}(t_{a})={\bf b}^{kf}(t_{a})+{\bf P}_{bb}^{f}{\bf M}^{% T}\left[{\bf M}{\bf P}_{bb}^{f}{\bf M}^{T}+{\bf R}_{yy}\right]^{-1}bold_b start_POSTSUPERSCRIPT italic_k italic_a end_POSTSUPERSCRIPT ( italic_t start_POSTSUBSCRIPT italic_a end_POSTSUBSCRIPT ) = bold_b start_POSTSUPERSCRIPT italic_k italic_f end_POSTSUPERSCRIPT ( italic_t start_POSTSUBSCRIPT italic_a end_POSTSUBSCRIPT ) + bold_P start_POSTSUBSCRIPT italic_b italic_b end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_f end_POSTSUPERSCRIPT bold_M start_POSTSUPERSCRIPT italic_T end_POSTSUPERSCRIPT [ bold_MP start_POSTSUBSCRIPT italic_b italic_b end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_f end_POSTSUPERSCRIPT bold_M start_POSTSUPERSCRIPT italic_T end_POSTSUPERSCRIPT + bold_R start_POSTSUBSCRIPT italic_y italic_y end_POSTSUBSCRIPT ] start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT
(𝐲ko(ta)𝐌𝐛kf(ta)),absentsuperscript𝐲𝑘𝑜subscript𝑡𝑎superscript𝐌𝐛𝑘𝑓subscript𝑡𝑎\displaystyle\cdot\left({\bf y}^{ko}(t_{a})-{\bf M}{\bf b}^{kf}(t_{a})\right)\,,⋅ ( bold_y start_POSTSUPERSCRIPT italic_k italic_o end_POSTSUPERSCRIPT ( italic_t start_POSTSUBSCRIPT italic_a end_POSTSUBSCRIPT ) - bold_Mb start_POSTSUPERSCRIPT italic_k italic_f end_POSTSUPERSCRIPT ( italic_t start_POSTSUBSCRIPT italic_a end_POSTSUBSCRIPT ) ) ,

with tasubscript𝑡𝑎t_{a}italic_t start_POSTSUBSCRIPT italic_a end_POSTSUBSCRIPT the analysis epoch and the superscript k𝑘kitalic_k referring to the kthsuperscript𝑘𝑡k^{th}italic_k start_POSTSUPERSCRIPT italic_t italic_h end_POSTSUPERSCRIPT realization within an ensemble chosen to be of size Nm=50subscript𝑁𝑚50N_{m}=50italic_N start_POSTSUBSCRIPT italic_m end_POSTSUBSCRIPT = 50. Writing as 𝐏zzfsuperscriptsubscript𝐏𝑧𝑧𝑓{\bf P}_{zz}^{f}bold_P start_POSTSUBSCRIPT italic_z italic_z end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_f end_POSTSUPERSCRIPT the forecast model covariance matrix for 𝐳𝐳{\bf z}bold_z, the second analysis (equation (20) of BGA17) is replaced here by {linenomath*}

k[1,Nm],for-all𝑘1subscript𝑁𝑚\displaystyle\forall k\in[1,N_{m}],\;∀ italic_k ∈ [ 1 , italic_N start_POSTSUBSCRIPT italic_m end_POSTSUBSCRIPT ] , (19)
𝐳ka(ta)=𝐳kf(ta)+𝐏zzf𝐆kT[𝐆k𝐏zzf𝐆kT+𝐑y˙y˙]1superscript𝐳𝑘𝑎subscript𝑡𝑎superscript𝐳𝑘𝑓subscript𝑡𝑎superscriptsubscript𝐏𝑧𝑧𝑓superscript𝐆𝑘𝑇superscriptdelimited-[]superscript𝐆𝑘superscriptsubscript𝐏𝑧𝑧𝑓superscript𝐆𝑘𝑇subscript𝐑˙𝑦˙𝑦1\displaystyle{\bf z}^{ka}(t_{a})={\bf z}^{kf}(t_{a})+{\bf P}_{zz}^{f}{\bf G}^{% kT}\left[{\bf G}^{k}{\bf P}_{zz}^{f}{\bf G}^{kT}+{\bf R}_{\dot{y}\dot{y}}% \right]^{-1}bold_z start_POSTSUPERSCRIPT italic_k italic_a end_POSTSUPERSCRIPT ( italic_t start_POSTSUBSCRIPT italic_a end_POSTSUBSCRIPT ) = bold_z start_POSTSUPERSCRIPT italic_k italic_f end_POSTSUPERSCRIPT ( italic_t start_POSTSUBSCRIPT italic_a end_POSTSUBSCRIPT ) + bold_P start_POSTSUBSCRIPT italic_z italic_z end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_f end_POSTSUPERSCRIPT bold_G start_POSTSUPERSCRIPT italic_k italic_T end_POSTSUPERSCRIPT [ bold_G start_POSTSUPERSCRIPT italic_k end_POSTSUPERSCRIPT bold_P start_POSTSUBSCRIPT italic_z italic_z end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_f end_POSTSUPERSCRIPT bold_G start_POSTSUPERSCRIPT italic_k italic_T end_POSTSUPERSCRIPT + bold_R start_POSTSUBSCRIPT over˙ start_ARG italic_y end_ARG over˙ start_ARG italic_y end_ARG end_POSTSUBSCRIPT ] start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT
(δ𝐲˙ko(ta)𝐆k𝐳kf(ta)),absent𝛿superscript˙𝐲𝑘𝑜subscript𝑡𝑎superscript𝐆𝑘superscript𝐳𝑘𝑓subscript𝑡𝑎\displaystyle\cdot\left(\delta\dot{\bf y}^{ko}(t_{a})-{\bf G}^{k}{\bf z}^{kf}(% t_{a})\right)\,,⋅ ( italic_δ over˙ start_ARG bold_y end_ARG start_POSTSUPERSCRIPT italic_k italic_o end_POSTSUPERSCRIPT ( italic_t start_POSTSUBSCRIPT italic_a end_POSTSUBSCRIPT ) - bold_G start_POSTSUPERSCRIPT italic_k end_POSTSUPERSCRIPT bold_z start_POSTSUPERSCRIPT italic_k italic_f end_POSTSUPERSCRIPT ( italic_t start_POSTSUBSCRIPT italic_a end_POSTSUBSCRIPT ) ) ,

where the new observation operator is 𝐆k=𝐌𝐇(𝐛ka)superscript𝐆𝑘𝐌𝐇superscript𝐛𝑘𝑎{\bf G}^{k}={\bf M}{\bf H}({\bf b}^{ka})bold_G start_POSTSUPERSCRIPT italic_k end_POSTSUPERSCRIPT = bold_MH ( bold_b start_POSTSUPERSCRIPT italic_k italic_a end_POSTSUPERSCRIPT ), with 𝐇𝐇{\bf H}bold_H as defined in BGA17. Here δ𝐲˙ko(ta)=𝐲˙ko(ta)𝐌𝐝kf(ta)𝛿superscript˙𝐲𝑘𝑜subscript𝑡𝑎superscript˙𝐲𝑘𝑜subscript𝑡𝑎superscript𝐌𝐝𝑘𝑓subscript𝑡𝑎\delta\dot{\bf y}^{ko}(t_{a})=\dot{\bf y}^{ko}(t_{a})-{\bf M}{\bf d}^{kf}(t_{a})italic_δ over˙ start_ARG bold_y end_ARG start_POSTSUPERSCRIPT italic_k italic_o end_POSTSUPERSCRIPT ( italic_t start_POSTSUBSCRIPT italic_a end_POSTSUBSCRIPT ) = over˙ start_ARG bold_y end_ARG start_POSTSUPERSCRIPT italic_k italic_o end_POSTSUPERSCRIPT ( italic_t start_POSTSUBSCRIPT italic_a end_POSTSUBSCRIPT ) - bold_Md start_POSTSUPERSCRIPT italic_k italic_f end_POSTSUPERSCRIPT ( italic_t start_POSTSUBSCRIPT italic_a end_POSTSUBSCRIPT ) are the direct SV observations corrected by the forecast contribution from diffusion to the radial induction equation. This latter is sought iteratively at each analysis step, as in BGA17. Note that we consider an ensemble of observations 𝐲osuperscript𝐲𝑜{\bf y}^{o}bold_y start_POSTSUPERSCRIPT italic_o end_POSTSUPERSCRIPT and 𝐲˙osuperscript˙𝐲𝑜\dot{\bf y}^{o}over˙ start_ARG bold_y end_ARG start_POSTSUPERSCRIPT italic_o end_POSTSUPERSCRIPT, which are perturbed by random noise according to respectively 𝐑yysubscript𝐑𝑦𝑦{\bf R}_{yy}bold_R start_POSTSUBSCRIPT italic_y italic_y end_POSTSUBSCRIPT and 𝐑y˙y˙subscript𝐑˙𝑦˙𝑦{\bf R}_{\dot{y}\dot{y}}bold_R start_POSTSUBSCRIPT over˙ start_ARG italic_y end_ARG over˙ start_ARG italic_y end_ARG end_POSTSUBSCRIPT. We recall that we consider in equations (1819) forecast covariance matrices 𝐏zzsubscript𝐏𝑧𝑧{\bf P}_{zz}bold_P start_POSTSUBSCRIPT italic_z italic_z end_POSTSUBSCRIPT and 𝐏bbfsuperscriptsubscript𝐏𝑏𝑏𝑓{\bf P}_{bb}^{f}bold_P start_POSTSUBSCRIPT italic_b italic_b end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_f end_POSTSUPERSCRIPT that are frozen throughout the re-analysis period. These are derived directly from the CED cross-covariances on 𝐛𝐛{\bf b}bold_b, 𝐮𝐮{\bf u}bold_u and 𝐞𝐞{\bf e}bold_e spherical harmonic coefficients, involving scaling prefactors obtained analytically from the stochastic model presented in §2.2.1 (see BGA17 for details). For comparison, Baerenzung et al.(2017)Baerenzung, Holschneider, Wicht, Sanchez, & Lesur employ a full implementation of the Ensemble Kalman filter (Evensen(2003)), i.e. they update the cross-covariances at each analysis step, requiring many more realizations to obtain well-conditioned matrices.

Finally, an extra complexity arises because the number of observation sites changes over time. Indeed, because of the selection criteria, the number of satellite data available may not always be sufficient to make a reliable VO estimate. Under these conditions the VO data point is considered to be absent: the associated elements of the data vector 𝐲o(t)superscript𝐲𝑜𝑡{\bf y}^{o}(t)bold_y start_POSTSUPERSCRIPT italic_o end_POSTSUPERSCRIPT ( italic_t ) at a given time t𝑡titalic_t are removed, together with the corresponding lines and columns of 𝐑yysubscript𝐑𝑦𝑦{\bf R}_{yy}bold_R start_POSTSUBSCRIPT italic_y italic_y end_POSTSUBSCRIPT, and the corresponding lines of the matrix 𝐌𝐌{\bf M}bold_M (and thus 𝐆𝐆{\bf G}bold_G). This procedure is performed during each analysis. Thus, the size P𝑃Pitalic_P of the data vector changes through time, reflecting the changing number of available satellite observations through time (see Figure 2).

Refer to caption

Figure 2: Time evolution of the number of SV data points (VOs in red, GOs in blue).

To summarise, in this study we work with predictions made by spherical harmonic coefficients that are projected in physical space, where they are adjusted during the analysis step according to the observations and the covariance matrices. As such, our algorithm is still based almost entirely on the spectral domain; only the analysis steps are performed in physical space, in order to match the observed magnetic field data. Notice that we corrected for two mistakes in the implementation of the algorithm by BGA17: a sign error in the background flow 𝐮^^𝐮\hat{\bf u}over^ start_ARG bold_u end_ARG, and off-diagonal elements of the covariance matrix for 𝐞𝐞{\bf e}bold_e were non intentionally ignored. Performing comparisons between re-analyses before and after correction, we found two consequences: a reduction of the dispersion within the ensemble of realizations, and an (almost stationary) shift in the analysed diffusion for some coefficients (including the axial dipole, see §3.1.2). This latter is almost entirely compensated by a shift in the analysed ersubscript𝑒𝑟e_{r}italic_e start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT, with minor impact on the recovered flow. Otherwise, the qualitative conclusions of BGA17 remain unaltered.

2.3 Posterior diagnostics

We now define several diagnostics used to evaluate the quality and the consistency of our results. We shall compare a quantity 𝒙𝒙{\bm{x}}bold_italic_x (MF, SV, subgrid error, diffusion… in the spatial or spectral domain) with observations 𝒙osuperscript𝒙𝑜{\bm{x}}^{o}bold_italic_x start_POSTSUPERSCRIPT italic_o end_POSTSUPERSCRIPT (when available), or with the same quantities 𝒙csuperscript𝒙𝑐{\bm{x}}^{c}bold_italic_x start_POSTSUPERSCRIPT italic_c end_POSTSUPERSCRIPT from the CHAOS-6 geomagnetic model (Finlay et al.(2016b)Finlay, Olsen, Kotsiaros, Gillet, & Tøffner-Clausen). We define its time average {linenomath*}

𝒙^=1tftititf𝒙(t)𝑑t,^𝒙1subscript𝑡𝑓subscript𝑡𝑖superscriptsubscriptsubscript𝑡𝑖subscript𝑡𝑓𝒙𝑡differential-d𝑡\displaystyle\hat{\bm{x}}=\frac{1}{t_{f}-t_{i}}\int_{t_{i}}^{t_{f}}{\bm{x}}(t)% dt\,,over^ start_ARG bold_italic_x end_ARG = divide start_ARG 1 end_ARG start_ARG italic_t start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT - italic_t start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT end_ARG ∫ start_POSTSUBSCRIPT italic_t start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_t start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT end_POSTSUPERSCRIPT bold_italic_x ( italic_t ) italic_d italic_t , (20)

with tisubscript𝑡𝑖t_{i}italic_t start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT and tfsubscript𝑡𝑓t_{f}italic_t start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT the initial and final epochs, its ensemble mean {linenomath*}

𝒙(t)=1Nmk=1Nm𝒙k(t),delimited-⟨⟩𝒙𝑡1subscript𝑁𝑚superscriptsubscript𝑘1subscript𝑁𝑚superscript𝒙𝑘𝑡\displaystyle\left<{\bm{x}}(t)\right>=\dfrac{1}{N_{m}}\sum_{k=1}^{N_{m}}{\bm{x% }}^{k}(t)\,,⟨ bold_italic_x ( italic_t ) ⟩ = divide start_ARG 1 end_ARG start_ARG italic_N start_POSTSUBSCRIPT italic_m end_POSTSUBSCRIPT end_ARG ∑ start_POSTSUBSCRIPT italic_k = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_N start_POSTSUBSCRIPT italic_m end_POSTSUBSCRIPT end_POSTSUPERSCRIPT bold_italic_x start_POSTSUPERSCRIPT italic_k end_POSTSUPERSCRIPT ( italic_t ) , (21)

the dispersion within the ensemble {linenomath*}

𝝈x(t)=1Nm1k=1Nm(𝒙k(t)𝒙(t))2,subscript𝝈𝑥𝑡1subscript𝑁𝑚1superscriptsubscript𝑘1subscript𝑁𝑚superscriptsuperscript𝒙𝑘𝑡delimited-⟨⟩𝒙𝑡2\displaystyle{\bm{\sigma}}_{x}(t)=\displaystyle\sqrt{\dfrac{1}{N_{m}-1}\sum_{k% =1}^{N_{m}}\left({\bm{x}}^{k}(t)-\left<{\bm{x}}(t)\right>\right)^{2}}\,,bold_italic_σ start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT ( italic_t ) = square-root start_ARG divide start_ARG 1 end_ARG start_ARG italic_N start_POSTSUBSCRIPT italic_m end_POSTSUBSCRIPT - 1 end_ARG ∑ start_POSTSUBSCRIPT italic_k = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_N start_POSTSUBSCRIPT italic_m end_POSTSUBSCRIPT end_POSTSUPERSCRIPT ( bold_italic_x start_POSTSUPERSCRIPT italic_k end_POSTSUPERSCRIPT ( italic_t ) - ⟨ bold_italic_x ( italic_t ) ⟩ ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG , (22)

and finally the bias between our ensemble mean model and the reference 𝒙csuperscript𝒙𝑐{\bm{x}}^{c}bold_italic_x start_POSTSUPERSCRIPT italic_c end_POSTSUPERSCRIPT, {linenomath*}

𝜹x(t)=𝒙c𝒙(t).subscript𝜹𝑥𝑡superscript𝒙𝑐delimited-⟨⟩𝒙𝑡\displaystyle{\bm{\delta}}_{x}(t)=\displaystyle{\bm{x}}^{c}-\left<{\bm{x}}(t)% \right>\,.bold_italic_δ start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT ( italic_t ) = bold_italic_x start_POSTSUPERSCRIPT italic_c end_POSTSUPERSCRIPT - ⟨ bold_italic_x ( italic_t ) ⟩ . (23)

We also define spatial power spectra of any magnetic trajectory 𝐛(t)𝐛𝑡{\bf b}(t)bold_b ( italic_t ) as {linenomath*}

b(n,t)=(n+1)(ac)2n+4m=0n[gnm(t)2+hnm(t)2],subscript𝑏𝑛𝑡𝑛1superscriptsuperscript𝑎direct-sum𝑐2𝑛4superscriptsubscript𝑚0𝑛delimited-[]superscriptsubscript𝑔𝑛𝑚superscript𝑡2superscriptsubscript𝑛𝑚superscript𝑡2\displaystyle{\cal R}_{b}(n,t)=(n+1)\left(\frac{a^{\oplus}}{c}\right)^{2n+4}% \sum_{m=0}^{n}\left[{g_{n}^{m}(t)}^{2}+{h_{n}^{m}(t)}^{2}\right]\,,caligraphic_R start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT ( italic_n , italic_t ) = ( italic_n + 1 ) ( divide start_ARG italic_a start_POSTSUPERSCRIPT ⊕ end_POSTSUPERSCRIPT end_ARG start_ARG italic_c end_ARG ) start_POSTSUPERSCRIPT 2 italic_n + 4 end_POSTSUPERSCRIPT ∑ start_POSTSUBSCRIPT italic_m = 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT [ italic_g start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_m end_POSTSUPERSCRIPT ( italic_t ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + italic_h start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_m end_POSTSUPERSCRIPT ( italic_t ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ] , (24)

with similar notations for 𝐛˙(t)˙𝐛𝑡\dot{\bf b}(t)over˙ start_ARG bold_b end_ARG ( italic_t ), 𝐝(t)𝐝𝑡{\bf d}(t)bold_d ( italic_t ) and 𝐞(t)𝐞𝑡{\bf e}(t)bold_e ( italic_t ). c=3485𝑐3485c=3485italic_c = 3485 km is the Earth’s core radius, and gnmsuperscriptsubscript𝑔𝑛𝑚g_{n}^{m}italic_g start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_m end_POSTSUPERSCRIPT and hnmsuperscriptsubscript𝑛𝑚h_{n}^{m}italic_h start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_m end_POSTSUPERSCRIPT are Schmidt semi-normalised spherical harmonic coefficients for the magnetic field at the Earth’s surface. Finally, the spatial power spectrum for core flow trajectories 𝐮𝐮{\bf u}bold_u writes {linenomath*}

𝒮(n,t)=n(n+1)2n+1m=0n[tcnm(t)2+tsnm(t)2+scnm(t)2+ssnm(t)2],𝒮𝑛𝑡𝑛𝑛12𝑛1superscriptsubscript𝑚0𝑛delimited-[]superscriptsubscriptsubscript𝑡𝑐𝑛𝑚superscript𝑡2superscriptsubscriptsubscript𝑡𝑠𝑛𝑚superscript𝑡2superscriptsubscriptsubscript𝑠𝑐𝑛𝑚superscript𝑡2superscriptsubscriptsubscript𝑠𝑠𝑛𝑚superscript𝑡2\displaystyle{\cal S}(n,t)=\frac{n(n+1)}{2n+1}\sum_{m=0}^{n}\left[{{t_{c}}_{n}% ^{m}(t)}^{2}+{{t_{s}}_{n}^{m}(t)}^{2}+{{s_{c}}_{n}^{m}(t)}^{2}+{{s_{s}}_{n}^{m% }(t)}^{2}\right]\,,caligraphic_S ( italic_n , italic_t ) = divide start_ARG italic_n ( italic_n + 1 ) end_ARG start_ARG 2 italic_n + 1 end_ARG ∑ start_POSTSUBSCRIPT italic_m = 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT [ italic_t start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_m end_POSTSUPERSCRIPT ( italic_t ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + italic_t start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_m end_POSTSUPERSCRIPT ( italic_t ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + italic_s start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_m end_POSTSUPERSCRIPT ( italic_t ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + italic_s start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_m end_POSTSUPERSCRIPT ( italic_t ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ] , (25)

with tc,snmsuperscriptsubscriptsubscript𝑡𝑐𝑠𝑛𝑚{{t_{c,s}}_{n}^{m}}italic_t start_POSTSUBSCRIPT italic_c , italic_s end_POSTSUBSCRIPT start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_m end_POSTSUPERSCRIPT and sc,snmsuperscriptsubscriptsubscript𝑠𝑐𝑠𝑛𝑚{{s_{c,s}}_{n}^{m}}italic_s start_POSTSUBSCRIPT italic_c , italic_s end_POSTSUBSCRIPT start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_m end_POSTSUPERSCRIPT Schmidt semi-normalised spherical harmonic coefficients for the toroidal and poloidal components of the flow. We also define the flow norm {linenomath*}

𝒩=n=1nun(n+1)2n+1m=0n[tcnm2+tsnm2+scnm2+ssnm2].𝒩superscriptsubscript𝑛1subscript𝑛𝑢𝑛𝑛12𝑛1superscriptsubscript𝑚0𝑛delimited-[]superscriptsuperscriptsubscriptsubscript𝑡𝑐𝑛𝑚2superscriptsuperscriptsubscriptsubscript𝑡𝑠𝑛𝑚2superscriptsuperscriptsubscriptsubscript𝑠𝑐𝑛𝑚2superscriptsuperscriptsubscriptsubscript𝑠𝑠𝑛𝑚2\displaystyle{\cal N}=\sum_{n=1}^{n_{u}}\frac{n(n+1)}{2n+1}\sum_{m=0}^{n}\left% [{{t_{c}}_{n}^{m}}^{2}+{{t_{s}}_{n}^{m}}^{2}+{{s_{c}}_{n}^{m}}^{2}+{{s_{s}}_{n% }^{m}}^{2}\right]\,.caligraphic_N = ∑ start_POSTSUBSCRIPT italic_n = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_n start_POSTSUBSCRIPT italic_u end_POSTSUBSCRIPT end_POSTSUPERSCRIPT divide start_ARG italic_n ( italic_n + 1 ) end_ARG start_ARG 2 italic_n + 1 end_ARG ∑ start_POSTSUBSCRIPT italic_m = 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT [ italic_t start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_m end_POSTSUPERSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + italic_t start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_m end_POSTSUPERSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + italic_s start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_m end_POSTSUPERSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + italic_s start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_m end_POSTSUPERSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ] . (26)

The above power spectra can be considered for the ensemble mean or the dispersion within the ensemble, in which case they are respectively noted <x>(n,t)subscriptexpectation𝑥𝑛𝑡{\cal R}_{<x>}(n,t)caligraphic_R start_POSTSUBSCRIPT < italic_x > end_POSTSUBSCRIPT ( italic_n , italic_t ) and δx(n,t)subscript𝛿𝑥𝑛𝑡{\cal R}_{\delta x}(n,t)caligraphic_R start_POSTSUBSCRIPT italic_δ italic_x end_POSTSUBSCRIPT ( italic_n , italic_t ). Additionally, all those quantities may be averaged in time and/or computed only at analysis periods. For example, the time-averaged spatial power spectrum of the dispersion of magnetic field solutions at analysis epochs is ^δba(n)superscriptsubscript^𝛿𝑏𝑎𝑛\hat{\cal R}_{\delta b}^{a}(n)over^ start_ARG caligraphic_R end_ARG start_POSTSUBSCRIPT italic_δ italic_b end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_a end_POSTSUPERSCRIPT ( italic_n ). The same convention as above holds for core flow spectra.

3 Results

We apply our algorithm to VO and GO magnetic field observations over a period spanning from ti=1996.92subscript𝑡𝑖1996.92t_{i}=1996.92italic_t start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT = 1996.92 to tf=2016.92subscript𝑡𝑓2016.92t_{f}=2016.92italic_t start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT = 2016.92. We recall that since we use satellite measurements from CHAMP and Swarm missions, VOs are available only over the period 2000–2010 and 2014–2017, whereas GOs are available over the whole time-span. Analysis are performed every Δta=4Δsuperscript𝑡𝑎4\Delta t^{a}=4roman_Δ italic_t start_POSTSUPERSCRIPT italic_a end_POSTSUPERSCRIPT = 4 months. The sequences of analyses and forecasts between 1997 and 2001 are used to warm-up the filter (see Figure 7 in BGA17), avoiding an increase in the ensemble spread over the first years of the targeted satellite era. This warm-up period is not considered below when interpreting the ensemble of inverted magnetic field and flow. We first describe predictions from our re-analysis for observations in the physical domain (§3.1.1), before we present the resulting magnetic model (§3.1.2), and insights on core flows over various time-scales (§3.2).

3.1 Geomagnetic field models

3.1.1 Predictions for GO and VO series

We compare in Figure 3 our series of SV forecasts and analysis with two examples of observation series (one VO and one GO), and with the predictions from CHAOS-6. The large spread of the SV forecasts is to be expected given the large uncertainties associated with subgrid errors and the large-scale flow (see BGA17). At both sites, the dispersion within the ensemble of SV trajectories encompasses most of the time the observations. Moreover the predictions from CHAOS-6 and from our ensemble of SV models are generally consistent. Our algorithm thus seems able to provide a coherent estimate of the SV probability density function (PDF) at the Earth’s surface and at satellite altitude. In addition, we highlight that even during the period 2010-2014 where no VO data are available, the trajectory of SV model, controlled by the stochastic prior and GO data only, remains reasonable, with a slight increase in the ensemble spread that always contains CHAOS-6. Note that our algorithm tends to drive the system toward low SV values (see the saw-tooth patterns in Figure 3). This feature is to be expected given our choice of the stochastic models for 𝒖Hsubscript𝒖𝐻{\bm{u}}_{H}bold_italic_u start_POSTSUBSCRIPT italic_H end_POSTSUBSCRIPT and ersubscript𝑒𝑟e_{r}italic_e start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT, which control the evolution of the SV. In the absence of data constraints, the process will drift back the ensemble average trajectories for 𝒖Hsubscript𝒖𝐻{\bm{u}}_{H}bold_italic_u start_POSTSUBSCRIPT italic_H end_POSTSUBSCRIPT and ersubscript𝑒𝑟e_{r}italic_e start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT towards the average dynamo state, which by construction is responsible for a weak SV. This is not a major drawback as soon as we analyse frequently enough, though it does limit the prediction capabilities of our tool (as discussed in BGA17).

Refer to caption
Refer to caption
Figure 3: SV time series for the three components (dBr/dt,dBθ/dt,dBϕ/dt)𝑑subscript𝐵𝑟𝑑𝑡𝑑subscript𝐵𝜃𝑑𝑡𝑑subscript𝐵italic-ϕ𝑑𝑡(dB_{r}/dt,dB_{\theta}/dt,dB_{\phi}/dt)( italic_d italic_B start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT / italic_d italic_t , italic_d italic_B start_POSTSUBSCRIPT italic_θ end_POSTSUBSCRIPT / italic_d italic_t , italic_d italic_B start_POSTSUBSCRIPT italic_ϕ end_POSTSUBSCRIPT / italic_d italic_t ), at one VO location {r=6671km,θ=90,ϕ=88,8}formulae-sequence𝑟6671𝑘𝑚formulae-sequence𝜃superscript90italic-ϕ88superscript8\left\{r=6671\;km,\theta=90^{\circ},\phi=88,8^{\circ}\right\}{ italic_r = 6671 italic_k italic_m , italic_θ = 90 start_POSTSUPERSCRIPT ∘ end_POSTSUPERSCRIPT , italic_ϕ = 88 , 8 start_POSTSUPERSCRIPT ∘ end_POSTSUPERSCRIPT } (top), and at Chambon-la-forêt {r=6366km,θ=42,ϕ=2}formulae-sequence𝑟6366𝑘𝑚formulae-sequence𝜃superscript42italic-ϕsuperscript2\left\{r=6366\;km,\theta=42^{\circ},\phi=2^{\circ}\right\}{ italic_r = 6366 italic_k italic_m , italic_θ = 42 start_POSTSUPERSCRIPT ∘ end_POSTSUPERSCRIPT , italic_ϕ = 2 start_POSTSUPERSCRIPT ∘ end_POSTSUPERSCRIPT } (bottom). SV observations are shown in black, CHAOS-6 predictions in green, predictions from our analysis in red. The shaded area correspond to ±σb˙plus-or-minussubscript𝜎˙𝑏\pm\sigma_{\dot{b}}± italic_σ start_POSTSUBSCRIPT over˙ start_ARG italic_b end_ARG end_POSTSUBSCRIPT, see equation (22).

We check in Figure 4 the accuracy with which our model fits MF and SV observations, with the histograms of the prediction errors (over all analyses) normalised to the observation errors, for the three components of the magnetic field. Concerning the MF, prediction errors are only weakly biased, excepted for Bθsubscript𝐵𝜃B_{\theta}italic_B start_POSTSUBSCRIPT italic_θ end_POSTSUBSCRIPT (normalised biases on the three components are μr=0.02subscript𝜇𝑟0.02\mu_{r}=-0.02italic_μ start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT = - 0.02, μθ=0.23subscript𝜇𝜃0.23\mu_{\theta}=-0.23italic_μ start_POSTSUBSCRIPT italic_θ end_POSTSUBSCRIPT = - 0.23 and μϕ=0.0subscript𝜇italic-ϕ0.0\mu_{\phi}=0.0italic_μ start_POSTSUBSCRIPT italic_ϕ end_POSTSUBSCRIPT = 0.0). The histograms of prediction errors are reasonably close to Gaussian for the three components with observation errors that appear to be under-estimated on average, in particular on Brsubscript𝐵𝑟B_{r}italic_B start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT (normalised r.m.s. errors on the three components are σr=2.18subscript𝜎𝑟2.18\sigma_{r}=2.18italic_σ start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT = 2.18, σθ=1.55subscript𝜎𝜃1.55\sigma_{\theta}=1.55italic_σ start_POSTSUBSCRIPT italic_θ end_POSTSUBSCRIPT = 1.55 and σϕ=1.63subscript𝜎italic-ϕ1.63\sigma_{\phi}=1.63italic_σ start_POSTSUBSCRIPT italic_ϕ end_POSTSUBSCRIPT = 1.63). The SV predictions errors are remarkably consistent with the a priori errors with small biases and standard deviation close to unity for the three components (μr=0.06subscript𝜇𝑟0.06\mu_{r}=-0.06italic_μ start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT = - 0.06, σr=1.01subscript𝜎𝑟1.01\sigma_{r}=1.01italic_σ start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT = 1.01; μθ=0.09subscript𝜇𝜃0.09\mu_{\theta}=-0.09italic_μ start_POSTSUBSCRIPT italic_θ end_POSTSUBSCRIPT = - 0.09, σθ=1.11subscript𝜎𝜃1.11\sigma_{\theta}=1.11italic_σ start_POSTSUBSCRIPT italic_θ end_POSTSUBSCRIPT = 1.11 and μϕ=0.03subscript𝜇italic-ϕ0.03\mu_{\phi}=0.03italic_μ start_POSTSUBSCRIPT italic_ϕ end_POSTSUBSCRIPT = 0.03, σϕ=1.14subscript𝜎italic-ϕ1.14\sigma_{\phi}=1.14italic_σ start_POSTSUBSCRIPT italic_ϕ end_POSTSUBSCRIPT = 1.14), even though the distributions appears more peaked than a Gaussian. The Kalman filter employed here implicitly assumes Gaussian distributed data errors. However, the above remark suggests that alternative treatments of data residuals may be worth considering in future studies (e.g. L1 or Huber norms, see Constable(1988); Farquharson & Oldenburg(1998)).

Refer to caption Refer to caption Refer to caption

Refer to caption Refer to caption Refer to caption

Figure 4: Top: Histograms of MF prediction errors δMFsubscript𝛿𝑀𝐹\delta_{MF}italic_δ start_POSTSUBSCRIPT italic_M italic_F end_POSTSUBSCRIPT (eq. 23), accumulated over all analysis epochs, normalised to the observation errors, for the components Brsubscript𝐵𝑟B_{r}italic_B start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT (left), Bθsubscript𝐵𝜃B_{\theta}italic_B start_POSTSUBSCRIPT italic_θ end_POSTSUBSCRIPT (middle) and Bϕsubscript𝐵italic-ϕB_{\phi}italic_B start_POSTSUBSCRIPT italic_ϕ end_POSTSUBSCRIPT (right). Superimposed in black are the Gaussian distribution fits obtained with the mean μ𝜇\muitalic_μ and the variance σ2superscript𝜎2\sigma^{2}italic_σ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT for each of the three distributions. Bottom: same histograms for the SV prediction errors δSVsubscript𝛿𝑆𝑉\delta_{SV}italic_δ start_POSTSUBSCRIPT italic_S italic_V end_POSTSUBSCRIPT.

3.1.2 Field models, and contributions to the SV

We now describe in more detail our MF and SV models. We present in Figure 5 MF and SV maps for our ensemble average model at the CMB truncated at spherical harmonic degree n=14𝑛14n=14italic_n = 14. Comparing it to a more traditional field model CHAOS-6, which is temporally regularised, the overall agreement is very good, indicating that our tool is indeed capable of producing reasonable field models. MF discrepancies to CHAOS-6 are relatively small, with peak to peak values less than 10%percent1010\%10 % of the total amplitude for a field truncated at degree 14. They are dominated by isotropically distributed, small length-scale patterns. As well as being dominated by small length-scales, the disagreements are larger for the SV, with peak to peak differences about 30%percent3030\%30 % of the total amplitude, which is to be expected given the blue SV spectrum at the CMB, meaning that small length scales dominate. Interestingly, the largest differences are localised under South America and the Indian Ocean, where the planetary gyre respectively detaches from and joins the equatorial belt (Pais & Jault(2008)) and where rapid time-dependence is observed (Finlay et al.(2016a)Finlay, Aubert, & Gillet).

Refer to caption Refer to caption

Refer to caption Refer to caption

Figure 5: Top: CMB maps of the ensemble average radial magnetic field 𝐛delimited-⟨⟩𝐛\left<{\bf b}\right>⟨ bold_b ⟩ (eq. 21) in 2017 (left: MF in mT; right: SV in μ𝜇\muitalic_μT/yr), as estimated with our algorithm. Bottom: MF (left) and SV (right) maps of the difference of our ensemble average field with CHAOS-6 (truncated at degree 14) at the CMB (with the same colorscales).

In Figure 6 we show the various contributions in our model to two SV spherical harmonic coefficient series. The dispersion within the ensemble of models is large enough to include time changes as estimated by CHAOS-6, with some exceptions during the high solar activity era, e.g. in 2002 for h66superscriptsubscript66h_{6}^{6}italic_h start_POSTSUBSCRIPT 6 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 6 end_POSTSUPERSCRIPT, and at the very end of the CHAOS-6 era (this latter possibly in link with the damping of SA towards end-points in the regularised field model). We notice a larger spread of the analysis for the axial dipole than for non-zonal coefficients of intermediate length-scale such as h66superscriptsubscript66h_{6}^{6}italic_h start_POSTSUBSCRIPT 6 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 6 end_POSTSUPERSCRIPT. This may be a consequence of the weaker constraint on zonal coefficients from surface observations (e.g. Kotsiaros & Olsen(2012)), although we only notice such behaviour for g10superscriptsubscript𝑔10g_{1}^{0}italic_g start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 0 end_POSTSUPERSCRIPT. An enhancement of the dispersion is noticeable between 2010 and 2014, displaying in the spectral domain the impact of the decreasing number of data during this era when no vector satellite data were available. Over 2001–2006, the ensemble average h66superscriptsubscript66h_{6}^{6}italic_h start_POSTSUBSCRIPT 6 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 6 end_POSTSUPERSCRIPT trajectory shows distinctive square shaped variations, probably partly related to variations in the number of data satisfying selection criteria during this interval of enhanced solar activity when only CHAMP data were available.

Refer to caption Refer to caption

Figure 6: SV spherical harmonic coefficient time series for g˙10superscriptsubscript˙𝑔10\dot{g}_{1}^{0}over˙ start_ARG italic_g end_ARG start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 0 end_POSTSUPERSCRIPT (left) and h˙66superscriptsubscript˙66\dot{h}_{6}^{6}over˙ start_ARG italic_h end_ARG start_POSTSUBSCRIPT 6 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 6 end_POSTSUPERSCRIPT (right). Predictions from our ensemble average model are shown in dark red (±2σb˙plus-or-minus2subscript𝜎˙𝑏\pm 2\sigma_{\dot{b}}± 2 italic_σ start_POSTSUBSCRIPT over˙ start_ARG italic_b end_ARG end_POSTSUBSCRIPT in red), and CHAOS-6 in black. Contributions from subgrid errors and diffusion extracted from our ensemble of realizations are superimposed in respectively blue and yellow (with dispersions ±1σdiffplus-or-minus1subscript𝜎𝑑𝑖𝑓𝑓\pm 1\sigma_{diff}± 1 italic_σ start_POSTSUBSCRIPT italic_d italic_i italic_f italic_f end_POSTSUBSCRIPT and ±1σerplus-or-minus1subscript𝜎𝑒𝑟\pm 1\sigma_{er}± 1 italic_σ start_POSTSUBSCRIPT italic_e italic_r end_POSTSUBSCRIPT in the corresponding colors).

Spatial spectra shown in Figure 7 summarise the characteristics of our model in the spectral domain. We find excellent agreement with CHAOS-6 for the main field and its secular variation, except at the small length scales of the SV (n>10𝑛10n>10italic_n > 10), which are more likely to be affected by the different data set chosen and by the different temporal kernel used (short time windows in our case against whole time-span for CHAOS-6). The ensemble spread gives a good approximation of the characteristic distance between our model and CHAOS-6. Diffusion and subgrid errors in the SV have approximately the same amplitude except for the dipole. The power stored in these two SV sources represents about 10101010 to 20%percent2020\%20 % of the total SV energy at all scales.

Even though the dispersion within the model predictions is large enough to encompass most of the MF and SV observations, the dispersion within the ensemble of realizations is lower, by a factor about 2.5, than the distance between the ensemble average model and CHAOS-6 for both the MF (at all length-scales) and the SV (towards small length-scales only). A complete account of SV errors from all subgrid interactions (see Baerenzung et al.(2017)Baerenzung, Holschneider, Wicht, Sanchez, & Lesur) may help reduce the above under-estimation. Our current estimate is nevertheless larger than that obtained for the COV-OBS.x1 model Gillet et al.(2015a)Gillet, Barrois, & Finlay. We suspect that the accumulation of data (assumed independent) during the construction of this latter field model involved too strong a decrease of the posterior error within the COV-OBS framework. The more consistent approach to error propagation developed here and presented in Figure 7 favours larger uncertainties on spherical harmonic coefficients during the satellite era.

Refer to caption
Refer to caption
Figure 7: Top: time averaged spatial power spectra at the Earth surface of the magnetic field of CHAOS-6 (^bcasuperscriptsubscript^superscript𝑏𝑐𝑎\hat{\cal R}_{b^{c}}^{a}over^ start_ARG caligraphic_R end_ARG start_POSTSUBSCRIPT italic_b start_POSTSUPERSCRIPT italic_c end_POSTSUPERSCRIPT end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_a end_POSTSUPERSCRIPT, eq. (24), in green), our estimate (^basuperscriptsubscript^delimited-⟨⟩𝑏𝑎\hat{\cal R}_{\left<b\right>}^{a}over^ start_ARG caligraphic_R end_ARG start_POSTSUBSCRIPT ⟨ italic_b ⟩ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_a end_POSTSUPERSCRIPT, in red), the difference between the two (^δbasuperscriptsubscript^𝛿𝑏𝑎\hat{\cal R}_{\delta b}^{a}over^ start_ARG caligraphic_R end_ARG start_POSTSUBSCRIPT italic_δ italic_b end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_a end_POSTSUPERSCRIPT, red thin line) and the dispersion within our ensemble of analyses (^σbasuperscriptsubscript^subscript𝜎𝑏𝑎\hat{\cal R}_{\sigma_{b}}^{a}over^ start_ARG caligraphic_R end_ARG start_POSTSUBSCRIPT italic_σ start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_a end_POSTSUPERSCRIPT, dotted line). Bottom: idem for the SV, superimposed with the spectra of the contributions from subgrid errors (blue) and from diffusion (yellow).

Overall, we are generally able to retrieve earlier well-established results. For instance the contribution from advection dominates (over diffusion) the axial dipole decay (Finlay et al.(2016b)Finlay, Olsen, Kotsiaros, Gillet, & Tøffner-Clausen; Barrois et al.(2017)Barrois, Gillet, & Aubert) and its fluctuations – even though our estimate for the contribution from diffusion to dg10/dt𝑑superscriptsubscript𝑔10𝑑𝑡dg_{1}^{0}/dtitalic_d italic_g start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 0 end_POSTSUPERSCRIPT / italic_d italic_t, shifted upward by a couple of nT/yr in comparison with the results of BGA17 (see §2.2.2), amounts to a relatively larger fraction over the latest years where the dipole decay tends to be weaker. The ensemble average SV originating from diffusion is presented in Figure 8 for 2017: the most significant contributions appear below Africa and Indonesia. The strongest diffusion appears linked to intense patches of up/down-wellings in the equatorial belt at the CMB (see Figure 8) and/or where strong gradient of 𝐁𝐁{\bf B}bold_B occur. This is a direct consequence of our estimation of diffusion through cross-covariances involving core surface velocity and magnetic fields (see BGA17 and Amit & Christensen(2008)). In the framework of our modelling, such diffusion patterns seem to be required by magnetic observations rather by the imposed prior cross-covariances (or if it is the case, it does not show up in the background state).

Refer to caption

Refer to caption

Figure 8: Magnetic diffusion at the CMB (top, color-scale in μ𝜇\muitalic_μT.y11{}^{-1}start_FLOATSUPERSCRIPT - 1 end_FLOATSUPERSCRIPT), and horizontal divergence h𝐮hsubscriptsubscript𝐮\nabla_{h}\cdot{\bf u}_{h}∇ start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT ⋅ bold_u start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT (bottom, color-scale in 103superscript10310^{-3}10 start_POSTSUPERSCRIPT - 3 end_POSTSUPERSCRIPT yrs11{}^{-1}start_FLOATSUPERSCRIPT - 1 end_FLOATSUPERSCRIPT) superimposed with passive tracers trajectories (black, tracer size scale in km/yr), for the ensemble average model in 2017. Core flow visualisations are performed using the tools provided at https://geodyn.univ-grenoble-alpes.fr/. The size of the tracers is proportional to the velocity field (see the legend). The initial positions of the tracers is random; each trajectory is advected by the velocity field for a fixed time; along each trajectory, the late (early) positions are darker (lighter).

3.2 Core flows solutions

Next, we study with more details the temporal information contained in our core flow solutions. The idea is to extract an average signal and a linear acceleration, together with the flow at different periods, to check if we witness any preferential frequency, or if the characteristics of the flow change with the period. To do so, we apply a least-squares regression to our core flow solution with a function of the form {linenomath*}

𝐮(t)=𝐀^+𝐀L(tt0)+k=111[𝐀kssin(2π(tt0)kT)\displaystyle{\bf u}(t)=\hat{\bf A}+{\bf A}_{L}(t-t_{0})+\displaystyle\sum_{k=% 1}^{11}\left[{\bf A}_{k}^{s}\sin\left(2\pi(t-t_{0})\dfrac{k}{T}\right)\right.bold_u ( italic_t ) = over^ start_ARG bold_A end_ARG + bold_A start_POSTSUBSCRIPT italic_L end_POSTSUBSCRIPT ( italic_t - italic_t start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ) + ∑ start_POSTSUBSCRIPT italic_k = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 11 end_POSTSUPERSCRIPT [ bold_A start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_s end_POSTSUPERSCRIPT roman_sin ( 2 italic_π ( italic_t - italic_t start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ) divide start_ARG italic_k end_ARG start_ARG italic_T end_ARG ) (27)
+𝐀kccos(2π(tt0)kT)],\displaystyle\left.+{\bf A}_{k}^{c}\cos\left(2\pi(t-t_{0})\dfrac{k}{T}\right)% \right]\,,+ bold_A start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_c end_POSTSUPERSCRIPT roman_cos ( 2 italic_π ( italic_t - italic_t start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ) divide start_ARG italic_k end_ARG start_ARG italic_T end_ARG ) ] ,

with t0=(ti+tf)/2=2008.92subscript𝑡0subscript𝑡𝑖subscript𝑡𝑓22008.92t_{0}=(t_{i}+t_{f})/2=2008.92italic_t start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = ( italic_t start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT + italic_t start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT ) / 2 = 2008.92 and T=tfti=16𝑇subscript𝑡𝑓subscript𝑡𝑖16T=t_{f}-t_{i}=16italic_T = italic_t start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT - italic_t start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT = 16 yrs. Vectors 𝐀^^𝐀\hat{\bf A}over^ start_ARG bold_A end_ARG, 𝐀Lsubscript𝐀𝐿{\bf A}_{L}bold_A start_POSTSUBSCRIPT italic_L end_POSTSUBSCRIPT, 𝐀kcsuperscriptsubscript𝐀𝑘𝑐{\bf A}_{k}^{c}bold_A start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_c end_POSTSUPERSCRIPT and 𝐀kssuperscriptsubscript𝐀𝑘𝑠{\bf A}_{k}^{s}bold_A start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_s end_POSTSUPERSCRIPT store respectively the spherical harmonic coefficients of the time average velocity, time average flow acceleration, and cosines and sines from periods 16161616 yrs (for k=1)k=1)italic_k = 1 ) to 1.451.451.451.45 yrs (for k=11𝑘11k=11italic_k = 11) – of course the longer periods are not well constrained given the short time-span considered here.

We show in Figure 9 the norm (26) of all flow constituents for the ensemble average solution. The flow is dominated by long periods, translating onto core surface motions the red SV temporal spectrum (see Gillet et al.(2015a)Gillet, Barrois, & Finlay; Lesur et al.(2017)Lesur, Wardinski, Baerenzung, & Holschneider). In comparison with a r.m.s. time average flow of 11.1 km/yr, the linear acceleration 𝐀Lsubscript𝐀𝐿{\bf A}_{L}bold_A start_POSTSUBSCRIPT italic_L end_POSTSUBSCRIPT corresponds, integrated over 16 yrs, to a r.m.s. flow increment of 6.6 km/yr.

Refer to caption

Figure 9: Core flow norm 𝒩𝒩{\cal N}caligraphic_N for all flow constituents that enter equation (27). The norm 𝒩𝒩\mathscr{N}script_N for the linear flow acceleration is obtained by integrating the linear trend over the 16 yrs.

3.2.1 Stationary motions, and flow model uncertainties

We show in Figure 10 core surface maps of the flow intensity and tracers trajectories for the ensemble average flow constituents 𝐀^^𝐀\hat{\bf A}over^ start_ARG bold_A end_ARG. We retrieve on the map for the time average flow classical features, such as the westward gyre offset towards the Atlantic Ocean found in many studies (e.g. Pais & Jault(2008); Gillet et al.(2015b)Gillet, Jault, & Finlay; Aubert(2014); Baerenzung et al.(2017)Baerenzung, Holschneider, Wicht, Sanchez, & Lesur), with a Pacific hemisphere that is on average much less energetic. The most energetic flow features are associated with (i) azimuthal motions in the equatorial belt below Africa, (ii) high latitudes azimuthal jets in the Pacific hemisphere and (iii) meridional circulations, poleward (resp. equatorward) around 90{}^{\circ}start_FLOATSUPERSCRIPT ∘ end_FLOATSUPERSCRIPTW (resp. 90{}^{\circ}start_FLOATSUPERSCRIPT ∘ end_FLOATSUPERSCRIPTE).

Refer to caption

Refer to caption Refer to caption

Refer to caption Refer to caption

Figure 10: Intensity maps at the CMB of the flow constituents 𝐀^^𝐀\hat{\bf A}over^ start_ARG bold_A end_ARG (in km/yr) for the ensemble average flow solution, superimposed with passive tracers trajectories (black). Top: Aitoff projection. Middle: North (right) and South (left) polar projections. Bottom: Aitoff projection for equatorially symmetric (left) and antisymmetric (right) components. The colorscale and tracer size scale are the same for all sub-figures.

Our solution is dominated by equatorially symmetric features (see Figure 10, bottom), as expected outside the tangent cylinder (or TC, the cylinder tangent to the inner core, whose axis coincides with to the rotation axis) when rotation forces dominates the momentum balance (e.g. Pais & Jault(2008)). Nevertheless, the symmetry may be locally broken. The most striking examples of this are anti-cyclonic circulations within the TC, retrieved in both the Northern and Southern hemispheres (Figure 10, middle). In contrast with polar vortices previously inferred from geomagnetic observations (Olson & Aurnou(1999); Amit & Olson(2006)), features we isolate here are off-set to one side of the polar caps (i.e. they contain an important m=1𝑚1m=1italic_m = 1 contribution). This is a common configuration for polar vortices found in the most up to date numerical simulations (Schaeffer et al.(2017)Schaeffer, Jault, Nataf, & Fournier), which show much variability through epochs.

We show in Figure 11 the time-average spatial power spectra for the ensemble average solution and for the dispersion within the ensemble of models. The former is comparable with the spectrum of the prior CED. The latter indicates that uncertainties, as measured by the ensemble spread, constitute a large fraction of the flow magnitude for degrees n10𝑛10n\geq 10italic_n ≥ 10. The oscillation in the power seen between odd and even degrees might be magnified by possibly too low subgrid error budget (see §3.1.2).

Refer to caption
Figure 11: Time averaged spatial power spectra for ensemble average core flows (𝒮^uasuperscriptsubscript^𝒮delimited-⟨⟩𝑢𝑎\hat{\cal S}_{\left<u\right>}^{a}over^ start_ARG caligraphic_S end_ARG start_POSTSUBSCRIPT ⟨ italic_u ⟩ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_a end_POSTSUPERSCRIPT, red thick line) and the spectra for the ensemble average of each realization (𝒮^ukadelimited-⟨⟩superscriptsubscript^𝒮𝑢𝑘𝑎\left<\hat{\cal S}_{uk}^{a}\right>⟨ over^ start_ARG caligraphic_S end_ARG start_POSTSUBSCRIPT italic_u italic_k end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_a end_POSTSUPERSCRIPT ⟩, yellow thick dotted line), eq. 25): obtained from the re-analysis of VO and GO data. Spectra for the corresponding dispersion within the ensembles of models are displayed in dotted lines. In green is shown the averaged spectrum for the prior CED.

3.2.2 On transient core surface motions

We now explore transient flow motions. We particularly focus on the amount of equatorial symmetry of our solutions inside and outside the TC, in order to detect if our model is sensitive to the specific geometry of the Earth’s core (does it hold a signature of the TC?). As for the time-average flow, the linear acceleration over the past 16 yrs is primarily symmetric with respect to the equator (see Figure 12). The largest contributions consist of accelerating circulations around the meridional, Eastern branch of the gyre. Associated with these time-changing eddies around the equatorward branch of the planetary gyre, an Eastward equatorial jet intensifies under the Western Pacific. This suggests an underlying dynamics more complex than a simple longitudinal shift of the planetary gyre.

Refer to caption

Refer to caption Refer to caption

Refer to caption Refer to caption

Figure 12: Same as Figure 10 for the flow constituent 𝐀Lsubscript𝐀𝐿{\bf A}_{L}bold_A start_POSTSUBSCRIPT italic_L end_POSTSUBSCRIPT (in km/yr22{}^{2}start_FLOATSUPERSCRIPT 2 end_FLOATSUPERSCRIPT).

Interestingly, our average solution does not show a major intensification of equatorially symmetric azimuthal jets at high latitudes in the Pacific hemisphere, as inferred by Livermore et al.(2017)Livermore, Hollerbach, & Finlay. Indeed, we see an increase of the Northern jet only, by about 67%percent6767\%67 % in average (the one σ𝜎\sigmaitalic_σ dispersion within the ensemble of flow realizations allowing for an increase up to 100%percent100100\%100 %). Although still an appreciable acceleration, it is significantly less than the factor of 3 found by Livermore et al.(2017)Livermore, Hollerbach, & Finlay. The disagreement is likely due to our global inversion (in opposition to their local model). The difference seems to be related with anti-symmetric circulations within the TC. One should keep in mind that in these high and low latitude areas, gradients of Brsubscript𝐵𝑟B_{r}italic_B start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT are much larger in the Northern hemisphere, meaning that the signature of any motions near the TC below the Southern Pacific are significantly weaker. As for the stationary constituent, the equatorial symmetry is not perfectly respected, and we retrieve the largest anti-symmetrical features within the TC, associated with polar jets.

We give in Figure 13 an example of one interannual flow constituent at the CMB for a period of 5.3 yrs. In this case, the most energetic flows are concentrated into non-axisymmetric azimuthal jets near the equator (already highlighted by Gillet et al.(2015b)Gillet, Jault, & Finlay; Finlay et al.(2016b)Finlay, Olsen, Kotsiaros, Gillet, & Tøffner-Clausen), and into localised circulations at mid and high latitudes. These are not confined to the Atlantic hemisphere: despite being less energetic on average, the Pacific hemisphere shows interesting interannual flow variations. At these sub-decadal periods, we have not detected any obvious propagation of non-zonal flow patterns, which might be interpreted as the signature of azimuthally propagating waves (as advocated for by Chulliat & Maus(2014); Chulliat et al.(2015)Chulliat, Alken, & Maus). The other periods display globally the same kind of features and no particular behaviour is found at any period. At these time-scales also show up less intense anti-symmetric features; the most significant shows up in the equatorial area (for instance under the Atlantic ocean and the Western Pacific), and towards high latitudes on the edge of the TC.

Refer to caption

Refer to caption Refer to caption

Refer to caption Refer to caption

Figure 13: Same as Figure 10 for the flow constituent 𝐀3csuperscriptsubscript𝐀3𝑐{\bf A}_{3}^{c}bold_A start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_c end_POSTSUPERSCRIPT (in km/yr).

Figure 14 summarises the amount of equatorial symmetry found in regions inside and outside the TC, for our core flow solutions at all periods. It appears almost independent of the considered period: outside the TC, it is within 90909090 to 95%percent9595\%95 % of the surface energy for all flow constituents of equation (27). The partition of energy between symmetric and anti-symmetric flow components is more balanced inside the TC where, depending on the considered time-scale, 55±15%absentplus-or-minus55percent15\approx 55\pm 15\%≈ 55 ± 15 % of the energy is contained in equatorially symmetric flows. This latter observation could be expected because the presence of the inner core is intended to partially break the equatorial symmetry However, it is remarkable that the algorithm appears accurate enough to detect a specific behaviour within the tiny areas covered by polar caps. Moreover, although our ensemble average model and the CED show very similar amounts of equatorial symmetry outside the TC (the value for the CED model is 95%percent9595\%95 % of symmetrical flows inside and outside TC), they differ significantly inside the TC (it is much less in the inverted flows). As a consequence, the larger proportion of equatorial antisymmetry inside the TC is driven by observations (against the prior information).

Refer to caption

Figure 14: Fraction of energy contained into the equatorial symmetric part of the flow, inside (blue line) and outside (yellow line) of the tangent cylinder (TC), for each of the flow constituent that enter equation (27). The total symmetric part of the flow is also displayed in green. The value for the CED dynamo used as a prior, is 0.950.950.950.95 both inside and outside the TC.

4 Summary and discussion

Following earlier strategies for geomagnetic field model reconstruction (e.g. Jackson et al.(2007)Jackson, Constable, Walker, & Parker; Lesur et al.(2010)Lesur, Wardinski, Asari, Minchev, & Mandea), and moving towards geomagnetic data assimilation (Aubert(2015); Gillet et al.(2015a)Gillet, Barrois, & Finlay; Baerenzung et al.(2017)Baerenzung, Holschneider, Wicht, Sanchez, & Lesur), we continue the work initiated in BGA17. We retain their idea of combining spatial information from numerical simulations of the geodynamo with temporal information implemented through stochastic equations, chosen to replicate the frequency spectrum of ground-based geomagnetic series. However, instead of considering spherical harmonic coefficients of the main field as data, here we have inverted observations (GOs and VOs) directly, at and above the Earth’s surface. In this respect we follow the studies by Beggan & Whaler(2009) and Whaler & Beggan(2015), although we account for subgrid processes (of great importance, as shown by BGA17 or Baerenzung et al.(2016)Baerenzung, Holschneider, & Lesur) and for surface magnetic diffusion. This avenue allows us to propose PDFs for the main field and its secular variation, as well as for the recovered core motions.

4.1 Geophysical insights

The MF models presented here are consistent both with observations and with the imposed dynamical prior. The model uncertainties, as suggested by the ensemble spread, are slightly less than the distance of the average model to CHAOS-6. We recover in our core flow solutions a westward gyre that circulates around the TC at high latitudes in the Pacific hemisphere, and flows closer to the equator in the Atlantic hemisphere. The largest contributions from magnetic diffusion are associated with up/down-wellings where the gyre meets the equatorial region (under Indonesia) and in the equatorial region below Africa. At all time-scales, the flow is predominantly symmetric with respect to the equator, except inside the TC where the situation is more balanced (contrary to our dynamo prior that is mostly symmetric everywhere).

The most intense time-average flow acceleration over the past 16 years is linked with evolving meanders around the equatorward branch of the gyre in the Eastern hemisphere, also associated with the appearance of an Eastward equatorial jet under the Western Pacific. We do find a decadal intensification of jets near the TC, although the magnitude of the acceleration we infer is lower than that estimated by Livermore et al.(2017)Livermore, Hollerbach, & Finlay with their reduced model. In our study, it is furthermore confined to the Northern hemisphere. This equatorial asymmetry may be interpreted as the signature of an ageostrophic acceleration, keeping in mind that main field gradients are weak in the Southern Pacific, implying a weaker constraint on flow motions there (see Figure 7 in Baerenzung et al.(2016)Baerenzung, Holschneider, & Lesur). However, because our prior does not show any particular bias in those areas, it is likely that those features are mostly driven by the data. On interannual periods, we find relatively energetic flow changes in both the Atlantic and the Pacific hemispheres, with both non-zonal equatorial jets and time-dependent mid-to-high latitudes eddies evident.

4.2 Future work

We currently lack a physical understanding for the features described above, whether it be through quasi-geostrophic flows (e.g. Labbé et al.(2015)Labbé, Jault, & Gillet), motions within a stratified layer (e.g. Buffett & Knezek(2017)), or any other interpretation through a reduced model. We also lack suitable long coverage by high quality satellite records to perform spectral analyses with a refined sampling in the frequency domain, which would allow us to isolate possible waves at interannual periods. Development of such reduced models, and their coupling with stochastic processes for modelling unresolved processes, will be an important next step in our ability to understand and predict geomagnetic field changes.

Meanwhile, our stochastic model itself could be improved; in particular it is desirable to avoid driving back the average trajectory towards an average dynamo simulation. This is indeed an unlikely state for the current era (say over decadal to centennial time-scales), which might be better represented by a re-analysis of for instance centennial motions from historical records (Jonkers et al.(2003)Jonkers, Jackson, & Murray). Furthermore, because of the short time-span covered today by satellite data, we found it challenging to derive well-conditioned matrices for VO uncertainties. This is a key-point for such data assimilation studies, which calls for further developments, e.g. through projections onto reduced basis in the data space. Alternatively, we may wish to co-estimate, together with the core state, time-dependent external fields. Although possible, this calls for a severe re-encoding of both the forecast and analysis steps, in order to integrate satellite measurements along the tracks.

The general philosophy of our work is to retrieve information on the state of the Earth’s core, and to provide realistic uncertainties on all state variables in a simple way. The encouraging magnetic models obtained with this approach render our algorithm suitable for deriving candidates to the International Geomagnetic Reference Field (Thébault et al.(2015)Thébault, Finlay, Beggan, Alken, Aubert, Barrois, Bertrand, Bondar, Boness, Brocco, et al.). Remaining in a stochastic framework, modifications of the forward model parametrisation – such as accounting for a background state closer to the flow responsible for the magnetic field over the past decades – may extend the prediction capability of our algorithm. However, targeting accurate field predictions one will have to resort to deterministic (i.e. dynamically based) equations for the core state.

5 Acknowledgements

We thank Julien Aubert for providing the Coupled-Earth dynamo series used to build the core state statistics, and Loic Huder for spotting two errors in the code at the origin of the results of BGA17. We also thank Julien Baerenzung and an anonymous referee for their useful comments that helped improve the quality of our manuscript. We would like to thank as well GFZ German Research Centre for Geoscience for providing access to the CHAMP MAG-L3 data and to ESA for providing access to the Swarm L1b MAG-L data. We also like to thank the staff of the geomagnetic observatories and INTERMAGNET for supplying high-quality observatory data. NG and OB were partially supported by the French Centre National d’Etudes Spatiales (CNES) for the study of Earth’s core dynamics in the context of the Swarm mission of ESA. ISTerre is part of Labex OSUG@2020 (ANR10 LABX56), which with the CNES also finance the phd grant of OB. Numerical computations were performed at the Froggy platform of the CIMENT infrastructure (https://ciment.ujf-grenoble.fr) supported by the Rhône-Alpes region (GRANT CPER07 13 CIRA), the OSUG@2020 Labex (reference ANR10 LABX56) and the Equip@Meso project (referenceANR-10-EQPX-29-01). MH and CF were supported by the Danish Council for Independent Research - Natural Sciences, Grant DFF-4002-00366.

References

  • [Amit & Christensen(2008)] Amit, H. & Christensen, U. R., 2008. Accounting for magnetic diffusion in core flow inversions from geomagnetic secular variation, Geophys. J. Int., 175(3), 913–924.
  • [Amit & Olson(2006)] Amit, H. & Olson, P., 2006. Time-average and time-dependent parts of core flow, Phys. Earth Planet. Int., 155, 120–139.
  • [Aubert(2014)] Aubert, J., 2014. Earth’s core internal dynamics 1840–2010 imaged by inverse geodynamo modelling, Geophys. J. Int., p. ggu064.
  • [Aubert(2015)] Aubert, J., 2015. Geomagnetic forecasts driven by thermal wind dynamics in the Earth’s core, Geophys. J. Int., 203(3), 1738–1751.
  • [Aubert et al.(2013)Aubert, Finlay, & Fournier] Aubert, J., Finlay, C. C., & Fournier, A., 2013. Bottom-up control of geomagnetic secular variation by the Earth’s inner core, Nature, 502(7470), 219–223.
  • [Baerenzung et al.(2016)Baerenzung, Holschneider, & Lesur] Baerenzung, J., Holschneider, M., & Lesur, V., 2016. The flow at the Earth’s core-mantle boundary under weak prior constraints, J. Geophys. Res.: Solid Earth, 121(3), 1343–1364.
  • [Baerenzung et al.(2017)Baerenzung, Holschneider, Wicht, Sanchez, & Lesur] Baerenzung, J., Holschneider, M., Wicht, J., Sanchez, S., & Lesur, V., 2017. Modeling and predicting the short term evolution of the Geomagnetic field, arXiv preprint arXiv:1710.06188.
  • [Barrois et al.(2017)Barrois, Gillet, & Aubert] Barrois, O., Gillet, N., & Aubert, J., 2017. Contributions to the geomagnetic secular variation from a reanalysis of core surface dynamics, Geophys. J. Int., 211(1), 50–68.
  • [Beggan & Whaler(2009)] Beggan, C. & Whaler, K., 2009. Forecasting change of the magnetic field using core surface flows and ensemble kalman filtering, Geophys. Res. Lett., 36(18).
  • [Beggan et al.(2009)Beggan, Whaler, & Macmillan] Beggan, C., Whaler, K., & Macmillan, S., 2009. Biased residuals of core flow models from satellite-derived ‘virtual observatories’, Geophysical Journal International, 177(2), 463–475.
  • [Buffett & Knezek(2017)] Buffett, B. A. & Knezek, N., 2017. Stochastic generation of MAC waves and implications for convection in Earth’s core, Geophys. J. Int..
  • [Chulliat & Maus(2014)] Chulliat, A. & Maus, S., 2014. Geomagnetic secular acceleration, jerks, and a localized standing wave at the core surface from 2000 to 2010, J. Geophys. Res.: Solid Earth, 119(3), 1531–1543.
  • [Chulliat & Olsen(2010)] Chulliat, A. & Olsen, N., 2010. Observation of magnetic diffusion in the Earth’s outer core from Magsat, Ørsted, and CHAMP data, J. Geophys. Res.: Solid Earth, 115(B5).
  • [Chulliat et al.(2015)Chulliat, Alken, & Maus] Chulliat, A., Alken, P., & Maus, S., 2015. Fast equatorial waves propagating at the top of the Earth’s core, Geophys. Res. Lett., 42(9), 3321–3329.
  • [Constable(1988)] Constable, C., 1988. Parameter estimation in non-Gaussian noise, Geophysical Journal International, 94(1), 131–142.
  • [Constable et al.(1993)Constable, Parker, & Stark] Constable, C. G., Parker, R. L., & Stark, P. B., 1993. Geomagnetic field models incorporating frozen-flux constraints, Geophys. J. Int., 113(2), 419–433.
  • [Evensen(2003)] Evensen, G., 2003. The ensemble kalman filter: Theoretical formulation and practical implementation, Ocean dynamics, 53(4), 343–367.
  • [Eymin & Hulot(2005)] Eymin, C. & Hulot, G., 2005. On core surface flows inferred from satellite magnetic data, Phys. Earth Planet. Int., 152(3), 200–220.
  • [Farquharson & Oldenburg(1998)] Farquharson, C. G. & Oldenburg, D. W., 1998. Non-linear inversion using general measures of data misfit and model structure, Geophys. J. Int., 134(1), 213–227.
  • [Finlay et al.(2016a)Finlay, Aubert, & Gillet] Finlay, C. C., Aubert, J., & Gillet, N., 2016a. Gyre-driven decay of the Earth’s magnetic dipole, Nature communications, 7, 10422.
  • [Finlay et al.(2016b)Finlay, Olsen, Kotsiaros, Gillet, & Tøffner-Clausen] Finlay, C. C., Olsen, N., Kotsiaros, S., Gillet, N., & Tøffner-Clausen, L., 2016b. Recent geomagnetic secular variation from Swarm, Earth, Planets and Space, 68(1), 1–18.
  • [Gillet et al.(2009)Gillet, Pais, & Jault] Gillet, N., Pais, M., & Jault, D., 2009. Ensemble inversion of time-dependent core flow models, Geochem. geophys. geosyst., 10(6).
  • [Gillet et al.(2015a)Gillet, Barrois, & Finlay] Gillet, N., Barrois, O., & Finlay, C. C., 2015a. Stochastic forecasting of the geomagnetic field from the COV-OBS. x1 geomagnetic field model, and candidate models for IGRF-12, Earth, Planets and Space, 67(1), 1–14.
  • [Gillet et al.(2015b)Gillet, Jault, & Finlay] Gillet, N., Jault, D., & Finlay, C., 2015b. Planetary gyre, time-dependent eddies, torsional waves, and equatorial jets at the Earth’s core surface, J. Geophys. Res.: Solid Earth, 120(6), 3991–4013.
  • [Holme(2015)] Holme, R., 2015. Large scale flow in the core, in Treatise in Geophysics, Core Dynamics, vol. 8, chap. 4, pp. 91–113, eds Olson, P. & Schubert, G., Elsevier.
  • [Jackson et al.(2007)Jackson, Constable, Walker, & Parker] Jackson, A., Constable, C., Walker, M., & Parker, R., 2007. Models of Earth’s main magnetic field incorporating flux and radial vorticity constraints, Geophys. J. Int., 171(1), 133–144.
  • [Jonkers et al.(2003)Jonkers, Jackson, & Murray] Jonkers, A. R., Jackson, A., & Murray, A., 2003. Four centuries of geomagnetic data from historical records, Reviews of Geophysics, 41(2).
  • [Kotsiaros & Olsen(2012)] Kotsiaros, S. & Olsen, N., 2012. The geomagnetic field gradient tensor, properties and parametrization in terms of spherical harmonics, Int. J. Geomath..
  • [Labbé et al.(2015)Labbé, Jault, & Gillet] Labbé, F., Jault, D., & Gillet, N., 2015. On magnetostrophic inertia-less waves in quasi-geostrophic models of planetary cores, Geophysical & Astrophysical Fluid Dynamics, 109(6), 587–610.
  • [Leopardi(2006)] Leopardi, P., 2006. A partition of the unit sphere into regions of equal area and small diameter, Electronic Transactions on Numerical Analysis, 25(12), 309–327.
  • [Lesur et al.(2010)Lesur, Wardinski, Asari, Minchev, & Mandea] Lesur, V., Wardinski, I., Asari, S., Minchev, B., & Mandea, M., 2010. Modelling the Earth’s core magnetic field under flow constraints, Earth, planets and space, 62(6), 503–516.
  • [Lesur et al.(2017)Lesur, Wardinski, Baerenzung, & Holschneider] Lesur, V., Wardinski, I., Baerenzung, J., & Holschneider, M., 2017. On the frequency spectra of the core magnetic field Gauss coefficients, Phys. Earth Planet. Int..
  • [Livermore et al.(2017)Livermore, Hollerbach, & Finlay] Livermore, P. W., Hollerbach, R., & Finlay, C. C., 2017. An accelerating high-latitude jet in Earth’s core, Nature Geoscience, 10(1), 62–68.
  • [Macmillan & Olsen(2013)] Macmillan, S. & Olsen, N., 2013. Observatory data and the Swarm mission, Earth, Planets and Space, 65(11), 15.
  • [Mandea & Olsen(2006)] Mandea, M. & Olsen, N., 2006. A new approach to directly determine the secular variation from magnetic satellite observations, Geophys. Res. Lett., 33(15).
  • [O’Brien et al.(1997)O’Brien, Constable, & Parker] O’Brien, M. S., Constable, C. G., & Parker, R. L., 1997. Frozen-flux modelling for epochs 1915 and 1980, Geophys. J. Int., 128(2), 434–450.
  • [Olsen(2002)] Olsen, N., 2002. A model of the geomagnetic field and its secular variation for epoch 2000 estimated from ørsted data, Geophysical Journal International, 149(2), 454–462.
  • [Olsen & Mandea(2007)] Olsen, N. & Mandea, M., 2007. Investigation of a secular variation impulse using satellite data: The 2003 geomagnetic jerk, Earth Planet. Sc. Lett., 255(1), 94–105.
  • [Olsen et al.(2010)Olsen, Glassmeier, & Jia] Olsen, N., Glassmeier, K.-H., & Jia, X., 2010. Separation of the magnetic field into external and internal parts, Space science reviews, 152(1-4), 135–157.
  • [Olsen et al.(2014)Olsen, Lühr, Finlay, Sabaka, Michaelis, Rauberg, & Tøffner-Clausen] Olsen, N., Lühr, H., Finlay, C. C., Sabaka, T. J., Michaelis, I., Rauberg, J., & Tøffner-Clausen, L., 2014. The CHAOS-4 geomagnetic field model, Geophys. J. Int., 197(2), 815–827.
  • [Olsen et al.(2015)Olsen, Hulot, Lesur, Finlay, Beggan, Chulliat, Sabaka, Floberghagen, Friis-Christensen, Haagmans, et al.] Olsen, N., Hulot, G., Lesur, V., Finlay, C. C., Beggan, C., Chulliat, A., Sabaka, T. J., Floberghagen, R., Friis-Christensen, E., Haagmans, R., et al., 2015. The swarm initial field model for the 2014 geomagnetic field, Geophysical Research Letters, 42(4), 1092–1098.
  • [Olson & Aurnou(1999)] Olson, P. & Aurnou, J., 1999. A polar vortex in the Earth’s core, Nature, 402(6758), 170–173.
  • [Pais & Jault(2008)] Pais, M. & Jault, D., 2008. Quasi-geostrophic flows responsible for the secular variation of the Earth’s magnetic field, Geophys. J. Int., 173(2), 421–443.
  • [Sabaka et al.(2004)Sabaka, Olsen, & Purucker] Sabaka, T. J., Olsen, N., & Purucker, M. E., 2004. Extending comprehensive models of the Earth’s magnetic field with Ørsted and CHAMP data, Geophys. J. Int., 159(2), 521–547.
  • [Sabaka et al.(2015)Sabaka, Olsen, Tyler, & Kuvshinov] Sabaka, T. J., Olsen, N., Tyler, R. H., & Kuvshinov, A., 2015. CM5, a pre-Swarm comprehensive geomagnetic field model derived from over 12 yr of CHAMP, Ørsted, SAC-C and observatory data, Geophys. J. Int., 200(3), 1596–1626.
  • [Schaeffer et al.(2017)Schaeffer, Jault, Nataf, & Fournier] Schaeffer, N., Jault, D., Nataf, H.-C., & Fournier, A., 2017. Turbulent geodynamo simulations: a leap towards Earth’s core, Geophys. J. Int., 211(1), 1–29.
  • [Thébault et al.(2015)Thébault, Finlay, Beggan, Alken, Aubert, Barrois, Bertrand, Bondar, Boness, Brocco, et al.] Thébault, E., Finlay, C. C., Beggan, C. D., Alken, P., Aubert, J., Barrois, O., Bertrand, F., Bondar, T., Boness, A., Brocco, L., et al., 2015. International geomagnetic reference field: the 12th generation, Earth, Planets and Space, 67(1), 1–19.
  • [Verboven & Hubert(2005)] Verboven, S. & Hubert, M., 2005. LIBRA: a MATLAB library for robust analysis, Chemometrics and intelligent laboratory systems, 75(2), 127–136.
  • [Wardinski & Lesur(2012)] Wardinski, I. & Lesur, V., 2012. An extended version of the C3FM geomagnetic field model: application of a continuous frozen-flux constraint, Geophys. J. Int., 189(3), 1409–1429.
  • [Whaler & Beggan(2015)] Whaler, K. & Beggan, C., 2015. Derivation and use of core surface flows for forecasting secular variation, J. Geophys. Res.: Solid Earth, 120(3), 1400–1414.
  • [Whaler et al.(2016)Whaler, Olsen, & Finlay] Whaler, K., Olsen, N., & Finlay, C., 2016. Decadal variability in core surface flows deduced from geomagnetic observatory monthly means, Geophysical Journal International, 207(1), 228–243.
IoCAEgkEkin0wQAfN9/cXPdheu6P33fBwB4ngcAcByHJpPJl+fn54mD3Gg0NrquXxeLRQAAwzAYj8cwTZPwPH9/sVg8PXweDAauqqr2cDjEer1GJBLBZDJBs9mE4zjwfZ85lAGg2+06hmGgXq+j3+/DsixYlgVN03a9Xu8jgCNCyIegIAgx13Vfd7vdu+FweG8YRkjXdWy329+dTgeSJD3ieZ7RNO0VAXAPwDEAO5VKndi2fWrb9jWl9Esul6PZbDY9Go1OZ7PZ9z/lyuD3OozU2wAAAABJRU5ErkJggg==" alt="[LOGO]">