Driscoll 2016

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 8

Geophysical Research Letters

RESEARCH LETTER Simulating 2 Ga of geodynamo history


10.1002/2016GL068858
Peter E. Driscoll1
Key Points:
1 Department of Terrestrial Magnetism, Carnegie Institution for Science, Washington, District of Columbia, USA
• Predicted signature of inner core
nucleation at ∼650 Ma is a transition
from weak-field nondipolar dynamo
to strong-field dipolar dynamo Abstract The paleomagnetic record indicates the geodynamo has been active over much of Earth history
• Neoproterozoic paleomagnetic
anomalies may be consistent with a with surprisingly little trend in paleointensity. Variability, however, is expected from models that predict a
transient weak-field geodynamo sharp increase in intensity following inner core nucleation (ICN) and implied by Neoproterozoic anomalies
• Dipolar dynamo in Mid-Proterozoic that hint at a highly variable field over several hundred million years. Here we demonstrate with a suite of
(1000–1700 Ma) similar to
Phanerozoic numerical dynamos driven by a new thermal evolution model that the geodynamo could have transitioned
from a multipolar to dipolar regime around 1.7 Ga, then to a weak-field dynamo around 1.0 Ga, and finally to
a strong-field dipole following ICN around 650 Ma that is maintained to the present day. The occurrence of
Supporting Information:
• Supporting Information S1 a weak-field geodynamo in the Neoproterozoic may be consistent with the observed anomalous apparent
polar wander paths and reversal behavior. Recovery to a dipolar geodynamo in the Phanerozoic could be
a signature of inner core nucleation. Index terms: 1507, 1560, and 1521.
Correspondence to:
P. E. Driscoll,
pdriscoll@ciw.edu

Citation:
1. Introduction
Driscoll, P. E. (2016), Simulating Global paleomagnetic observations indicate that the geodynamo has been active over at least the past 3.5 Ga
2 Ga of geodynamo history,
Geophys. Res. Lett., 43, 5680–5687, [Usui et al., 2009; Biggin et al., 2011] and possibly even 4.2 Ga [Tarduno et al., 2015]. The paleomagnetic record
doi:10.1002/2016GL068858. includes hundreds of polarity reversals [Lowrie and Kent, 2004], at least 3 and possibly 12 superchrons (periods
where the field lingered in a single polarity for tens of million years) [Pavlov and Gallet, 2005; Driscoll and
Received 4 FEB 2016 Evans, 2016] and hundreds of paleointensity measurements [Tauxe and Yamazaki, 2007; Biggin et al., 2015].
Accepted 10 MAY 2016 Suspiciously absent are significant long-term trends in paleointensity that are expected from the thermal
Accepted article online 16 MAY 2016 evolution of the core, nucleation of the inner core, changes in mantle convection associated with superconti-
Published online 6 JUN 2016
nental cycles, and superchron episodicity.
Numerical simulations of the core have been able to produce “Earth-like” magnetic fields similar to the
present-day field [e.g., Christensen et al., 2010], but limited progress has been made in connecting models
to observations of the ancient geodynamo. One feature, in particular, that has remained enigmatic is the
expected paleomagnetic signature of inner core nucleation (ICN) [e.g., Labrosse and Macouin, 2003; Aubert
et al., 2009]. Recently, Biggin et al. [2015] proposed that a statistical increase in paleointensity and variability in
the Mesoproterozoic (1–1.5 Ga) could be associated with ICN, although the reliability of some data has been
questioned [Smirnov et al., 2016]. More importantly, the global paleointensity record remains sparse through
the Neoproterozoic (0.54–1.0 Ga), which could be due to a scarcity of preservation or low reliability criteria
of samples.
In addition to providing a unique window into the dynamics of the core, paleomagnetic observations are cen-
tral to reconstructing the tectonic history of the planet via cratonic apparent polar wander paths (APWPs).
These reconstructions require that the time-averaged paleomagnetic field conforms to a global axial dipole
(GAD) field, which is supported by most, but not all, observations over the last 2 Ga. Successful plate recon-
structions through the Phanerozoic (0–541 Ma) and Mid-Proterozoic (1000–2219 Ma) have revealed a global
history of crustal dynamics, episodes of true polar wander, and an ∼600 Myr supercontinent episodicity [e.g.,
Torsvik et al., 2012; Evans, 2013]. Paleomagnetic data have also played an important role in interpretations of
low latitude glacial events, so-called “snowball Earth” scenarios [Hoffman and Schrag, 2002].
Despite these advances, however, several important discrepancies in Neoproterozoic paleogeography persist,
including the history of Rodinia [Evans, 2013; Li et al., 2013] and a growing number of anomalous poles in
the Ediacaran [Kirschvink et al., 1997; Abrajevitch and Van der Voo, 2010; Halls et al., 2015; Bono and Tarduno,
©2016. American Geophysical Union. 2015; Klein et al., 2015; Bazhenov et al., 2016] and Cryogenian [Maloof et al., 2006; Swanson-Hysell et al., 2012].
All Rights Reserved. Anomalous APWPs have been interpreted as either (i) rapid plate motions much faster than any occurring

DRISCOLL SIMULATING GEODYNAMO HISTORY 5680


Geophysical Research Letters 10.1002/2016GL068858

Figure 1. (a) Temperature evolution of mantle and core of preferred thermal evolution model and (b) predicted dipole
moment derived from a dynamo scaling law applied to the preferred core energy evolution. Also in Figure 1b):
inner core radius (red curve, right scale) and virtual dipole moment (VDM) paleointensity data from the Absolute
Paleointensity (PINT) Database (circles) [Biggin et al., 2009]. VDM data is the average within 100 Myr long bins with
symbol size in logarithmic proportion to the number of data in each bin, error bars denote standard deviation
of average within each bin, and bins with a single data point are shown as dots without error bars.

today, (ii) true polar wander (TPW) events where the entire lithosphere rotates to accommodate a change in
global moment of inertia, or (iii) the absence of a stable GAD field, including the possibility of an equatorial
dipole or “hyper” frequent polarity reversals [Abrajevitch and Van der Voo, 2010; Halls et al., 2015; Bazhenov
et al., 2016].
Geodynamo behavior during the Neoproterozoic, in particular, is of growing interest in light of recent revi-
sions to the timing of inner core nucleation that place it around 500–800 Ma [Driscoll and Bercovici, 2014;
Davies, 2015; Labrosse, 2015]. These estimates were motivated in part by upward revisions of the electrical
and thermal conductivity of high pressure-temperature iron [Pozzo et al., 2012; de Koker et al., 2012; Gomi et al.,
2013; Seagle et al., 2013], which imply that an adiabatic heat flow of 12–15 TW is required to keep the core well
mixed. This new constraint, dubbed the “New Core Paradox” [Olson, 2013], implies a rapid core cooling rate to
maintain a thermal dynamo prior to ICN. This “paradox” then joined the “thermal catastrophe” of the mantle
as two significant problems in the thermal history of the Earth. Driscoll and Bercovici [2014] demonstrated
that an adjustment to the present-day energy budget, increasing core heat flow to 15 TW and decreasing the
mantle secular cooling rate to 11 TW, produced models that successfully avoided both the core paradox and
the mantle “catastrophe” and predicted an inner core only 500–800 Myr old.
Previous models of long-term geodynamo evolution have considered the systematic influence of inner core
size on the stability of the dipole [e.g., Roberts and Glatzmaier, 2001; Coe and Glatzmaier, 2006] or the implica-
tions of a magnetic field scaling law applied to a thermal history [Aubert et al., 2009]. Here for the first time we
employ a hybrid approach using the energetic and inner core evolution implied by a whole-planet thermal
history to drive a suite of first principles magnetohydrodynamic geodynamo simulations, producing a 2 Gyr
span of paleodynamo behavior. Our results provide a set of predictions for the morphology and variability of
the paleomagnetic field and a guide for the applicability of the GAD assumption. We compare the results with
observations and offer a new prediction for the paleomagnetic signature of inner core nucleation.

2. Thermal Evolution Model


The numerical thermal evolution model is similar to Driscoll and Bercovici [2014], with two minor improve-
ments: (1) continental crust growth with radiogenic enrichment and (2) compositional dependence of the
core liquidus (see supporting information for details). These additions have a small effect on the core thermal
history, producing a slightly earlier inner core nucleation (ICN) time of 652 Ma (Figure 1) compared to 550 Ma
of Driscoll and Bercovici [2014]. Our preferred thermal evolution model produces a present-day state (TUM =
1648 K, TCMB = 4000 K, Ric = 1227 km, 𝜏ic = 652 Ma, Qsurf = 40.5 TW, QCMB = 11.5 TW, Qrad,man = 12.7 TW, Qrad,CC =
7.2 TW, Qrad,core = 1.5 TW, Urey ratio Qrad,man ∕Qsurf = 0.31, kc = 100 Wm−1 K−1 ) consistent with preferred

DRISCOLL SIMULATING GEODYNAMO HISTORY 5681


Geophysical Research Letters 10.1002/2016GL068858

Figure 2. Summary of numerical dynamo evolution over 0–2 Ga. (a) Continuous core power history from thermal
evolution model used to power the numerical dynamo models. (b) Time-averaged numerical dynamo dipole moment
scaled assuming present-day model is 50 ZAm2 . (c) Time-averaged ratio of axial to total dipole intensity at CMB.
(d) Time-averaged dipole latitude. In Figures 2b–2d circles denote individual dynamo models and are filled in
proportion to present-day inner core size, smoothed curve connecting models is added for clarity, and error bars
indicate one standard deviation of the mean.

present-day values (Figures 1a and 1b). Total core buoyancy power is always positive (Figure 2a) implying core
convection at all times. This nominal thermal evolution avoids both a thermal catastrophe in the mantle and
premature shutdown of the dynamo.
Before introducing the numerical dynamo models it is instructive to compute the dipole moment evolution
predicted from a magnetic scaling law applied to the preferred thermal evolution model, similar to what has
been done previously [Aubert et al., 2009; Driscoll and Bercovici, 2014]. The dynamo scaling law, which relates
core buoyancy flux to dipole moment assuming a dipole dominant dynamo [Olson and Christensen, 2006],
predicts a dipole moment decrease to 20 ZAm2 prior to ICN, followed by growth to a modern-day dipole
moment of 50 ZAm2 assisted by the compositional buoyancy of inner core growth (Figure 1b). We note that
the paleointensity averaged over the last 5 Myr of ∼50 ZAm2 [Wang et al., 2015] is less than the present-day
value of ∼80 ZAm2 . Also shown in Figure 1b are virtual dipole moment (VDM) data from the PINT paleointen-
sity database [Biggin et al., 2009] averaged in 100 Myr bins, with symbol size in proportion to number of data
in each bin.
The shape of the magnetic moment history in Figure 1b is typical of this type of calculation with a steady
decrease prior to ICN and a sharp jump after ICN: a trend that is not readily observed in the paleointensity
record (overplotted). However, numerical dynamos have demonstrated that a decrease in convective power
can cause a transition from a dipolar strong-field state, the assumed state for applying the magnetic scaling
law, to a metastable weak-field dynamo state and finally to a subdynamo state [e.g., Aubert and Wicht, 2004;
Morin and Dormy, 2009]. Therefore, the applicability of the dipole scaling law becomes questionable as the
model approaches the strong-field to weak-field regime, as may occur when the driving power is low. Next
we turn to first principles dynamo models driven by the preferred core power evolution (Figure 2a) to predict
the intensity and morphology of the paleomagnetic field over the last 2 Gyr.

3. Evolving Dynamo Models


A total of 25 dynamo models are computed at discrete times over 0–2 Ga with roughly equal 0.1 Gyr
spacing. Each model is driven by the core power predicted by the preferred thermal evolution model at that

DRISCOLL SIMULATING GEODYNAMO HISTORY 5682


Geophysical Research Letters 10.1002/2016GL068858

time (Figure 2a). Shorter time spacing between models is used in the Neoproterozoic and around ICN to
sample the time variability in greater detail. Each model is run at least 10 magnetic diffusion times, with an
average run length of ∼2 Myr. Starting at 2 Ga, the total core buoyancy power (thermal + compositional)
decreases monotonically prior to ICN and is boosted by IC growth during the Phanerozoic (Figure 2a). The
fixed numerical dynamo control parameters (Ra = 5 × 106 , E = 3 × 10−4 , Pr = 1, Pm = 10) are chosen such that
they produce an Earth-like magnetic field with present-day inner/outer core radius ratio (𝜖 = 0.35) and driving
power but are modest enough to allow for a number of long simulations to produce steady state magnetic
field properties.
The thermal evolution model is scaled to provide boundary conditions to drive the numerical dynamo model
by three steps (also see supporting information). First, an Earth-like dynamo is obtained at the preset-day
state, where the present-day inner core size and thermal/compositional power ratio are determined from
the preferred thermal evolution model. Our present-day Earth-like dynamo model has a magnetic field com-
pliance ratio of 𝜒 2 = 8.2, consistent with the Earth-like range defined by Christensen et al. [2010]. Second, a
fractional scaling is used to compute the initial (t0 ) conditions of the dynamo model. This fractional scaling
assumes the ratio of power P0 , at initial time t0 , to the present-day power P∗ of the thermal evolution model
is equal to the same ratio in the dynamo model, i.e., P0 ∕P∗ = P0′ ∕P′∗ , where ′ refers to the nondimensional
dynamo model power. This provides the dimensionless dynamo model power P0′ at t0 .
Third, a normalized scaling is used to compute dimensionless dynamo model power from the thermal evo-
lution model for all intervening times. This normalized scaling assumes that the normalized power of the
thermal evolution model is the same as the dynamo model, i.e., P(t) = P′ (t), where the normalized power

is P(t) = (P(t) − P∗ )∕(P0 − P∗ ) and similarly for P . This scaling is used to compute the time evolution of the
core-mantle boundary (CMB) and inner core boundary (ICB) heat flows Q′c and Q′i by replacing P′ in the
above with Q′c and Q′i . The internal heating buoyancy power H′ is then computed by energy conservation:
H′ (t) = Q′c (t) − Q′i (t).
The spherical coordinate system of the dynamo model requires that a finite inner core be present in all models,
including those prior to ICN. Therefore, as an approximation for all models prior to ICN we impose zero inner
core boundary (ICB) buoyancy flux, a free-slip ICB velocity condition, and a sufficiently small inner/outer core
radius ratio of 𝜖 = 0.1 so that the steady state magnetic field at the CMB does not depend on the presence of
the IC. Volumetrically 𝜖 = 0.1 corresponds to a dimensionless inner core that is less than 1% of the present-day
inner core volume (using 𝜖 ∗ = 0.35). To test this approximation, we computed a series of models over a range
of radius ratios (𝜖 = 0.05–0.1) driven by the same total power. We find that 𝜖<0.1 has a minor effect on the
time-averaged dynamo solution (in terms of energies and dipolarity), confirming that our approach is valid
for models prior to ICN (also see supporting information).
Other dynamo boundary conditions are as follows. The top velocity boundary condition is always no slip. The
bottom boundary condition changes from free slip prior to ICN to no slip after ICN. For all models a uniform
heat flux is imposed on the ICB, a uniform heat source/sink is imposed in the outer core, and the CMB
temperature is isothermal.

4. Results
The 25 dynamo models computed over 0 to 2 Ga can be categorized into three dynamo regimes (Figure 2):
(i) a strong-field dynamo where the dipole field dominates and is axially aligned (0–0.6 Ga and 1.0–1.7 Ga),
(ii) a multipolar strong-field dynamo where nondipole components are similar in magnitude to the dipole and
the dipole is not always axially aligned (1.7–2.0 Ga), and (iii) a weak-field dynamo where magnetic field inten-
sity is low and the dipole is generally not axially aligned (0.6–1.0 Ga). The strong-field dynamos with magnetic
Reynolds numbers Rm = 500–700 maintain a steady state balance between Coriolis and Lorentz forces such
that the flow and magnetic field are fully developed and saturated. In contrast, the weak-field dynamos with
Rm = 300–450 are kinetically dominated with low-amplitude, small-scale magnetic fields passively controlled
by the large-scale flow. The occurrence of each regime is determined by a combination of the total driving
power (Figure 2a) and depth of driving force. As generally found in previous dynamo studies, bottom-driven
dynamos promote strong axial dipoles as in regime (i), high-power top-driven dynamos promote strong
multipolar fields as in regime (ii), and low-power dynamos promote weak nonaxial fields as in regime (iii)
[Kutzner and Christensen, 2004; Aubert and Wicht, 2004; Christensen and Aubert, 2006; Driscoll and Olson, 2009;
Morin and Dormy, 2009].

DRISCOLL SIMULATING GEODYNAMO HISTORY 5683


Geophysical Research Letters 10.1002/2016GL068858

Figure 3. Nondimensional (nd) dynamo power vs scaled dipole moment. Symbols are the same as in Figure 2. Arrows
denote time evolution prior to ICN (red) and after ICN (blue). Dynamo regimes (i)–(iii) are also indicated.

Throughout the Phanerozoic the models occupy regime (i) with strong axially aligned dipole fields (Figures 2b
and 2c). Prior to ICN in the Neoproterozoic (0.6–1.0 Ga) the models occupy regime (iii): weak-field dynamos
with nonaxial time-averaged dipole fields (Figures 2b and 2c) and near-equatorial time-averaged dipole
latitudes (Figure 2d). Interestingly, during the Mid-Proterozoic (1.0–1.7 Ga) the thermally powered models are
nearly indiscriminate from the Phanerozoic, occupying dynamo regime (i) with strong axially aligned dipole
fields. Finally, Early Proterozoic models (1.7–2.0 Ga) tend toward dynamo regime (ii): strong multipolar fields
with midlatitude time-averaged dipole tilts (Figures 2c and 2d) where the dipole field is less favored due to
an increase in small scale dynamics and magnetic field generation. Although the Early Proterozoic models
produce similar dipole intensity and average latitude as those in the Neoproterozoic, the dynamo regimes are
significantly different: the Early Proterozoic (1.7–2.0 Ga) dynamos occupy the strong-field multipolar regime
(ii), while the Neoproterozoic (0.6–1.0 Ga) dynamos occupy the weak-field equatorial dipole regime (iii). The
kinetic and magnetic energy of the strong-field state is about 60% larger, promoting small-scale convection
and magnetic induction.
Surprisingly, models in the two strong-field dynamo periods (0–0.6 Ga and 1–1.7 Ga) produce similar dipole
strength (Figure 2b) despite the factor of ∼2 difference in driving power (Figure 2a). This similarity may be
coincidental, as the strong-field state prior to ICN is a top-driven dynamo, which has a different sensitivity to
driving power than the bottom driven model after ICN, or may indicate that strong-field dynamos saturate at
a particular field intensity independent of driving mode.
The evolution through these three distinct dynamo regimes can be further understood by examining dipole
moment as a function of total driving power (Figure 3). The dynamos begin at 2 Ga with medium power
(∼11 in nondimensional (nd) units) in a multipolar and weak dipole state until 1.7 Ga, at which point they
evolve steeply to the upper left due to decreasing power and a strengthening of the dipole from 1.05 to
1.60 Ga. As the power decreases further (to ∼8 (nd)) the models transition steeply down to a weak-field regime
(iii) occupied for 0.65–1 Ga. Finally, following ICN at 652 Ma, the models transition back up to a strong-field
dipolar regime (i), maintained to present day.
Time-averaged magnetic field maps illustrate the evolution of the morphology of the magnetic field through
the four dynamo regimes (Figure 4). Each map is a time-averaged sample of the field over one to two dipole
decay times (one dipole decay time is ∼50 kyr), which does not necessarily capture the full time variability of
the nondipolar models. At 2.0 Ga the strong-field multipolar model produces a somewhat dipolar time aver-
aged field, but with significant nondipolar components and modest field intensity (Figure 4d). The magnetic
field at 1.5 Ga and present day is strongly dipolar, axially aligned, and high intensity (Figures 4c and 4a). In
its weakest state, just prior to ICN at 0.67 Ga, the field is highly nonaxial, with significant multipolar compo-
nents and low amplitude. Although the field is highly time variable, the time average segment in Figure 4b is
representative of the equatorial and small-scale nature of the field through this period.

DRISCOLL SIMULATING GEODYNAMO HISTORY 5684


Geophysical Research Letters 10.1002/2016GL068858

Figure 4. Time average surface radial magnetic fields from models at (a) present day, (b) just prior to ICN at 0.67 Ga,
(c) 1.5 Ga, and (d) 2.0 Ga. Length of time averages are 45.5 kyr (Figure 4a), 82.7 kyr (Figure 4b), 62.0 kyr (Figure 4c), and
49.0 kyr (Figure 4d), assuming a dipole decay time of 50 kyr. Note Figures 4b and 4d are segments of highly time
variable models. Color scales show dimensionless model field intensity.

Importantly, the magnetic field evolution predicted by these numerical models (Figure 2b) differs from the
magnetic scaling law prediction (Figure 1b). The scaling law predicts a direct correspondence between field
intensity and buoyancy power, while the dynamo models predict a bimodal behavior, with sharp transitions
from strong field to weak field as the models cross dynamo regime boundaries. The paleodynamo scaling of
Aubert et al. [2009] accounted for some changes in driving mode and geometry, producing a slightly flatter
paleointensity prediction than the magnetic scaling law alone, but found no weak-field states. The occurrence
of weak-field states depends on the core power history, which is significantly different here (Figure 2a) than
in Aubert et al. [2009] due to differences in thermal history models, heat budgets, and core conductivities.

5. Implications
In light of these results, it is helpful to organize the implications for the last 2 Ga of geomagnetic history
into three periods: (1) the Phanerozoic (0–541 Ma) where a roughly continuous paleomagnetic and paleo-
geographic history is well known, (2) the Neoproterozoic (541–1000 Ma) where the discontinuous geologic
record contains periods of abnormal magnetic field behavior, and discrepant and rapid plate motions, and (3)
the Mid-Proterozoic (1000–2000 Ma), similar to the so-called “Boring Billion,” where the record reflects a more
Phanerozoic-like behavior with stable magnetic field, modest plate speeds, continuous paleogeography, and
fewer low-paleolatitude glaciations.
The main prediction of our simulations is that the geodynamo slipped into a weak-field state around 1 Ga due
to a secular decrease in thermal buoyancy power available to drive core flow and remained in this weak-field
state until sometime after inner core nucleation (∼650 Ma). Henceforth, inner core growth provided a boost
of compositional buoyancy power that pushed the geodynamo back to a strong-field state where it remained
throughout the Phanerozoic.
A weak-field geodynamo in the Neoproterozoic implies that some, perhaps many, APWPs during that time
were influenced by a highly time variable and at times significantly nonaxial dipole field. Observations during
such a weak field could produce rapid APWPs as the field oscillated between an axial and equatorial dipole,
with frequent apparent “polarity” reversals. More broadly, these results imply that the paleomagnetic signa-
ture of inner core nucleation may not have been a simple boost in dipole moment intensity as predicted by
magnetic scaling laws but rather a complicated morphological transition from a weak-field to strong-field
state around the Proterozoic-Phanerozoic transition. This transition would have been accompanied by a
switch from a transient and time variable nonaxial dipole to a stable axial dipole with occasional reversals.
The predicted stability of the axial dipole during the Mesoproterozoic and Phanerozoic is consistent with
the well-constrained global paleogeography during those times [Pisarevsky et al., 2014; Torsvik et al., 2012].

DRISCOLL SIMULATING GEODYNAMO HISTORY 5685


Geophysical Research Letters 10.1002/2016GL068858

The multipolar behavior found in the Early Proterozoic (1.7–2.0 Ga) may also be supported by observations
of rapid APWPs [Mitchell et al., 2010] and deserves further investigation.
Equatorial dipoles have been found in weak-field numerical dynamo regimes [Kuang and Bloxham, 1999;
Aubert and Wicht, 2004] but have not previously been applied to the geodynamo. Even if the Neoproterozoic
geodynamo contained a significant equatorial dipole component, it is not obvious that such a field would
be preserved in a time-averaged magnetic remanence because the equatorial dipole is expected to precess
westward with the global drift of core flow on an ∼1 kyr time scale. Obtaining a stable time average of a pre-
cessing equatorial dipole superimposed with a weak axial dipole likely requires many precession times, on the
order of a dipole decay time or longer [e.g., Davies and Constable, 2014]. Therefore, it is reasonable to expect
a magnetic remanence acquired over a few kiloyears to preserve some equatorial components of the field
(as in Figure 4b), if present, and for concurrent global measurements to predict discordant APWPs. In reality,
APWPs are a complex superposition of fluctuations in the local geomagnetic pole and plate motions, and
untangling the two may require a hybrid modeling scheme where both geologic measurements and dynam-
ical models are interpreted in conjunction.
Our modeling approach using a secular thermal evolution limits applicability of the dynamo results to the
steady state behavior of the geodynamo and does not include the time variability expected from mantle
dynamics. For example, the ∼200 Myr quasi-periodic fluctuation in polarity reversal frequency during the
Phanerozoic is likely driven by fluctuations in the core cooling rate or its morphology at the CMB as determined
by the lower mantle [e.g., Olson et al., 2015]. Time variability of secular variation and paleointensity could also
be temporarily enhanced by fluctuations in CMB heat flux. However, our results do imply that the geodynamo
operated close to the reversing-nonreversing regime boundary during the Mid-Proterozoic and Phanerozoic,
so that mantle imposed fluctuations in CMB heat flux could have driven the geodynamo temporarily through
frequently reversing and nonreversing states [Driscoll and Olson, 2009].
Several caveats deserve mention. First, the thermal evolution model that provides the power and inner core
growth history to drive the dynamo models is not unique and is subject to significant uncertainties. Revisions
to core material properties and the global heat balance could adjust the timing of ICN and the dynamo transi-
tions predicted here by ±100 Myr. Additional buoyancy sources, for example, through the exsolution of a light
element [e.g., O’Rourke and Stevenson, 2016], could operate prior to ICN, and future work in this area is needed.
Second, the dynamo model suffers from the normal numerical difficulties, namely, that the model operates
at parameters far from Earth like. To address this, we chose a present-day dynamo model that produces an
Earth-like magnetic field and rely on the driving forces from the thermal evolution model to generate numer-
ical paleodynamo predictions. The sensitivity of more realistic dynamo simulations to time variable control
parameters should be addressed in future studies. Third, our Neoproterozoic weak-field dynamos with non-
axial dipoles do not produce the virtual geomagnetic pole longitudinal clustering proposed by Abrajevitch
and Van der Voo [2010]. Such longitudinal asymmetries in a nonaxial weak-field dipole could be produced
by a heterogeneous cooling pattern at the CMB, as has been proposed for several Phanerozoic reversals
[Love, 2000; Kutzner and Christensen, 2004], and is worthy of future study.

Acknowledgments References
We are grateful for comments by
J. Aubert, D. Evans, and an anonymous Abrajevitch, A., and R. Van der Voo (2010), Incompatible Ediacaran paleomagnetic directions suggest an equatorial geomagnetic dipole
reviewer that greatly improved this hypothesis, Earth Planet. Sci. Lett., 293(1), 164–170.
manuscript. Numerical simulations Aubert, J., and J. Wicht (2004), Axial vs. equatorial dipolar dynamo models with implications for planetary magnetic fields, Earth Planet. Sci.
were computed with the code MagIC Lett., 221(1), 409–419.
developed by J. Wicht and performed Aubert, J., S. Labrosse, and C. Poitou (2009), Modeling the paleo-evolution of the geodynamo, Geophys. J. Int., 179(3), 1414–1428,
at the Memex High Performance doi:10.1111/j.1365-246X.2009.04,361.x.
Computing Cluster at the Carnegie Bazhenov, M. L., N. M. Levashova, J. G. Meert, I. V. Golovanova, K. N. Danukalov, and N. M. Fedorova (2016), Late Ediacaran
Institution for Science. The data magnetostratigraphy of Baltica: Evidence for magnetic field hyperactivity?, Earth Planet. Sci. Lett., 435, 124–135.
described in this paper are available Biggin, A., G. Strik, and C. Langereis (2009), The intensity of the geomagnetic field in the late-Archaean: New measurements and an analysis
upon request (pdriscoll@ciw.edu). of the updated IAGA palaeointensity database, Earth Planets Space, 61(1), 9–22.
Biggin, A., E. Piispa, L. Pesonen, R. Holme, G. Paterson, T. Veikkolainen, and L. Tauxe (2015), Palaeomagnetic field intensity variations suggest
mesoproterozoic inner-core nucleation, Nature, 526(7572), 245–248.
Biggin, A. J., M. J. de Wit, C. G. Langereis, T. E. Zegers, S. Voûte, M. J. Dekkers, and K. Drost (2011), Palaeomagnetism of archaean rocks of the
onverwacht group, barberton greenstone belt (southern Africa): Evidence for a stable and potentially reversing geomagnetic field at
ca. 3.5 Ga, Earth Planet. Sci. Lett., 302(3), 314–328.
Bono, R. K., and J. A. Tarduno (2015), A stable Ediacaran Earth recorded by single silicate crystals of the ca. 565 Ma Sept-Iles intrusion,
Geology, 43(2), 131–134.
Christensen, U., and J. Aubert (2006), Scaling properties of convection-driven dynamos in rotating spherical shells and application to
planetary magnetic fields, Geophys. J. Int., 166(1), 97–114, doi:10.1111/j.1365-246X.2006.03,009.x.

DRISCOLL SIMULATING GEODYNAMO HISTORY 5686


Geophysical Research Letters 10.1002/2016GL068858

Christensen, U., J. Aubert, and G. Hulot (2010), Conditions for Earth-like geodynamo models, Earth Planet. Sci. Lett., 296, 487–496.
Coe, R., and G. Glatzmaier (2006), Symmetry and stability of the geomagnetic field, Geophys. Res. Lett., 33, L21311,
doi:10.1029/2006GL027903.
Davies, C. J. (2015), Cooling history of Earth’s core with high thermal conductivity, Phys. Earth Planet. Inter., 247, 65–79.
Davies, C. J., and C. G. Constable (2014), Insights from geodynamo simulations into long-term geomagnetic field behaviour, Earth Planet.
Sci. Lett., 404, 238–249.
de Koker, N., G. Steinle-Neumann, and V. Vlċek (2012), Electrical resistivity and thermal conductivity of liquid Fe alloys at high P and T, and
heat flux in Earth’s core, Proc. Natl. Acad. Sci., 109(11), 4070–4073.
Driscoll, P., and D. Bercovici (2014), On the thermal and magnetic histories of Earth and Venus: Influences of melting, radioactivity,
and conductivity, Phys. Earth Planet. Inter., 236, 36–51.
Driscoll, P., and P. Olson (2009), Effects of buoyancy and rotation on the polarity reversal frequency of gravitationally driven numerical
dynamos, Geophys. J. Int., 178(3), 1337–1350, doi:10.1111/j.1365-246X.2009.04,234.x.
Driscoll, P. E., and D. A. Evans (2016), Frequency of proterozoic geomagnetic superchrons, Earth Planet. Sci. Lett., 437, 9–14,
doi:10.1016/j.epsl.2015.12.035.
Evans, D. A. (2013), Reconstructing pre-pangean supercontinents, Geol. Soc. Am. Bull., 125(11–12), 1735–1751.
Gomi, H., K. Ohta, K. Hirose, S. Labrosse, R. Caracas, M. J. Verstraete, and J. W. Hernlund (2013), The high conductivity of iron and thermal
evolution of the Earth’s core, Phys. Earth Planet. Inter., 224, 88–103.
Halls, H. C., A. Lovette, M. Hamilton, and U. Söderlund (2015), A paleomagnetic and U–Pb geochronology study of the western end of the
Grenville dyke swarm: Rapid changes in paleomagnetic field direction at ca. 585 Ma related to polarity reversals?, Precambrian Res., 257,
137–166.
Hoffman, P. F., and D. P. Schrag (2002), The snowball earth hypothesis: Testing the limits of global change, Terra Nova, 14(3), 129–155.
Kirschvink, J. L., R. L. Ripperdan, and D. A. Evans (1997), Evidence for a large-scale reorganization of early cambrian continental masses
by inertial interchange true polar wander, Science, 277(5325), 541–545.
Klein, R., J. Salminen, and S. Mertanen (2015), Baltica during the Ediacaran and Cambrian: A paleomagnetic study of Hailuoto sediments in
Finland, Precambrian Res., 267, 94–105.
Kuang, W., and J. Bloxham (1999), Numerical modeling of magnetohydrodynamic convection in a rapidly rotating spherical shell: Weak and
strong field dynamo action, J. Computat. Phys., 153(1), 51–81.
Kutzner, C., and U. Christensen (2004), Simulated geomagnetic reversals and preferred virtual geomagnetic pole paths, Geophys. J. Int.,
157(3), 1105–1118.
Labrosse, S. (2015), Thermal evolution of the core with a high thermal conductivity, Phys. Earth Planet. Inter., 247, 36–55.
Labrosse, S., and M. Macouin (2003), The inner core and the geodynamo, C. R. Geosci., 335(1), 37–50.
Li, Z.-X., D. A. Evans, and G. P. Halverson (2013), Neoproterozoic glaciations in a revised global palaeogeography from the breakup of Rodinia
to the assembly of Gondwanaland, Sediment. Geol., 294, 219–232.
Love, J. (2000), Statistical assessment of preferred transitional VGP longitudes based on palaeomagnetic lava data, Geophys. J. Int., 140(1),
211–221.
Lowrie, W., and D. Kent (2004), Geomagnetic polarity timescales and reversal frequency regimes, in Chapman Conference on Timescales of
the Internal Geomagnetic Field, Geophys. Monogr. Ser., vol. 145, pp. 117–129, AGU, Washington, D. C., doi:10.1029/145GM09.
Maloof, A. C., G. P. Halverson, J. L. Kirschvink, D. P. Schrag, B. P. Weiss, and P. F. Hoffman (2006), Combined paleomagnetic, isotopic, and
stratigraphic evidence for true polar wander from the Neoproterozoic Akademikerbreen Group, Svalbard, Norway, Geol. Soc. Am. Bull.,
118(9–10), 1099–1124.
Mitchell, R. N., P. F. Hoffman, and D. A. Evans (2010), Coronation loop resurrected: Oscillatory apparent polar wander of Orosirian
(2.05–1.8 Ga) paleomagnetic poles from Slave craton, Precambrian Res., 179(1), 121–134.
Morin, V., and E. Dormy (2009), The dynamo bifurcation in rotating spherical shells, Int. J. Mod. Phys. B, 23(28–29), 5467–5482.
Olson, P. (2013), The new core paradox, Science, 342(6157), 431–432, doi:10.1126/science.1243477.
Olson, P., and U. Christensen (2006), Dipole moment scaling for convection-driven planetary dynamos, Earth Planet. Sci. Lett., 250(3–4),
561–571.
Olson, P., R. Deguen, M. L. Rudolph, and S. Zhong (2015), Core evolution driven by mantle global circulation, Phys. Earth Planet. Inter., 243,
44–55.
O’Rourke, J. G., and D. J. Stevenson (2016), Powering Earth’s dynamo with magnesium precipitation from the core, Nature, 529(7586),
387–389.
Pavlov, V., and Y. Gallet (2005), A third superchron during the early Paleozoic, Episodes, 28(2), 78–84.
Pisarevsky, S. A., S.-Å. Elming, L. J. Pesonen, and Z.-X. Li (2014), Mesoproterozoic paleogeography: Supercontinent and beyond,
Precambrian Res., 244, 207–225.
Pozzo, M., C. Davies, D. Gubbins, and D. Alfè (2012), Thermal and electrical conductivity of iron at Earth’s core conditions, Nature, 485(7398),
355–358.
Roberts, P., and G. Glatzmaier (2001), The geodynamo, past, present and future, Geophys. Astrophys. Fluid Dynam., 94(1), 47–84.
Seagle, C. T., E. Cottrell, Y. Fei, D. R. Hummer, and V. B. Prakapenka (2013), Electrical and thermal transport properties of iron and iron-silicon
alloy at high pressure, Geophys. Res. Lett., 40, 5377–5381, doi:10.1002/2013GL057930.
Smirnov, A. V., J. A. Tarduno, E. V. Kulakov, S. A. McEnroe, and R. K. Bono (2016), Paleointensity, core thermal conductivity and the unknown
age of the inner core, Geophys. J. Int., 205, 1190–1195, doi:10.1093/gji/ggw080.
Swanson-Hysell, N. L., A. C. Maloof, J. L. Kirschvink, D. A. Evans, G. P. Halverson, and M. T. Hurtgen (2012), Constraints on neoproterozoic
paleogeography and paleozoic orogenesis from paleomagnetic records of the bitter springs formation, amadeus basin, central australia,
Am. J. Sci., 312(8), 817–884.
Tarduno, J. A., R. D. Cottrell, W. J. Davis, F. Nimmo, and R. K. Bono (2015), A Hadean to Paleoarchean geodynamo recorded by single zircon
crystals, Science, 349(6247), 521–524.
Tauxe, L., and T. Yamazaki (2007), Paleointensities, Treatise on Geophysics, vol. 5, Geomagnetism, chap. 13, pp. 509–563, Elsevier, Amsterdam.
Torsvik, T. H., et al. (2012), Phanerozoic polar wander, palaeogeography and dynamics, Earth Sci. Rev., 114, 325–368.
Usui, Y., J. A. Tarduno, M. Watkeys, A. Hofmann, and R. D. Cottrell (2009), Evidence for a 3.45-billion-year-old magnetic remanence: Hints of
an ancient geodynamo from conglomerates of South Africa, Geochem. Geophys. Geosyst., 10, Q09Z07, doi:10.1029/2009GC002496.
Wang, H., D. V. Kent, and P. Rochette (2015), Weaker axially dipolar time-averaged paleomagnetic field based on multidomain-corrected
paleointensities from Galapagos lavas, Proc. Natl. Acad. Sci., 112(49), 15,036–15,041.

DRISCOLL SIMULATING GEODYNAMO HISTORY 5687

You might also like