Sei sulla pagina 1di 28

EA47CH03_Nakagawa ARjats.

cls April 20, 2019 10:38

Annual Review of Earth and Planetary Sciences


Dynamics in the Uppermost
Lower Mantle: Insights into the
Deep Mantle Water Cycle
Annu. Rev. Earth Planet. Sci. 2019.47:41-66. Downloaded from www.annualreviews.org

Based on the Numerical


Access provided by 186.183.193.242 on 05/03/20. For personal use only.

Modeling of Subducted Slabs


and Global-Scale Mantle
Dynamics
Takashi Nakagawa1,2 and Tomoeki Nakakuki3
1
Department of Mathematical Science and Technology, Japan Agency for Marine-Earth
Science and Technology, Yokohama 236-0001, Japan; email: ntakashi@jamstec.go.jp
2
Current affiliation: Department of Earth Sciences, The University of Hong Kong,
Hong Kong; email: ntakashi@hku.hk
3
Department of Earth and Planetary Systems Science, Hiroshima University,
Higashi-Hiroshima 739-8526, japan; email: nakakuki@hiroshima-u.ac.jp

Annu. Rev. Earth Planet. Sci. 2019. 47:41–66 Keywords


First published as a Review in Advance on
mantle convection, mantle layering, stagnant slab, phase transitions, water
December 10, 2018

The Annual Review of Earth and Planetary Sciences is Abstract


online at earth.annualreviews.org
In this review, we address the current status of numerical modeling of the
https://doi.org/10.1146/annurev-earth-053018-
mantle transition zone and uppermost lower mantle, focusing on the hy-
060305
dration mechanism in these areas. The main points are as follows: (a) Slab
Copyright © 2019 by Annual Reviews.
stagnation and penetration may play significant roles in transporting the
All rights reserved
water in the whole mantle, and (b) a huge amount of water could be ab-
sorbed into the deep mantle to preserve the surface seawater over the ge-
ologic timescale. However, for further understanding of water circulation
in the deep planetary interior, more mineral physics investigations are re-
quired to reveal the mechanism of water absorption in the lower mantle
and thermochemical interaction across the core–mantle boundary region,

41
EA47CH03_Nakagawa ARjats.cls April 20, 2019 10:38

which can provide information on material properties to the geodynamics community. Moreover,
future investigations should focus on determining the amount of water in the early planetary
interior, as suggested by the planetary formation theory of rocky planets. Moreover, the supplying
mechanism of water during planetary formation and its evolution caused by plate tectonics are still
essential issues because, in geodynamics modeling, a huge amount of water seems to be required
to preserve the surface seawater in the present day and to not be dependent on an initial amount
of water in Earth’s system.

 Slab stagnation and penetration of the hydrous lithosphere are essential for understanding
the global-scale material circulation.
 Thermal feedback caused by water-dependent viscosity is a main driving mechanism of
water absorption in the mantle transition zone and uppermost lower mantle.
Annu. Rev. Earth Planet. Sci. 2019.47:41-66. Downloaded from www.annualreviews.org

 The hydrous state in the early rocky planets remains to be determined from cosmo- and
geochemistry and planetary formation theory.
Access provided by 186.183.193.242 on 05/03/20. For personal use only.

 Volatile cycles in the deep planetary interior may affect the evolution of the surface
environment.

1. INTRODUCTION
Earth’s mantle is approximately 2,890 km thick and is divided into two regions (the upper and
lower mantle) at a depth of 660 km. This depth is recognized as a seismic discontinuity (e.g.,
Deuss 2009). This type of seismic discontinuity is interpreted as the phase transition between the
spinel and the postspinel, which may have a negative Clapeyron slope (Ito & Takahashi 1989);
that is, it is currently recognized as a bridgmanite transition. In addition to this phase transition, a
high-viscosity jump has been indicated by geophysical (gravity anomalies) (King 1995, Thoraval &
Richards 1997) and environmental (sea-level change) observations (e.g., Nakada & Okuno 2016),
as well as the joint inversion of both geoid and sea-level change (Kaufmann & Lambeck 2000,
Mitrovica & Forte 1997), which indicates that the lower mantle is 5–100 times more viscous than
the upper mantle. The viscosity structure in the mantle has also been proposed to exhibit a viscosity
hill in the mid-lower mantle, based on geoid data analysis (Rudolph et al. 2015) and the melting
temperature measured at lower mantle conditions (Deng & Lee 2017).
Recently, global tomographic images, which focus on the mantle transition zone to the up-
permost lower mantle, have indicated that various complicated dynamics can be observed in the
mantle transition zone and uppermost lower mantle; these dynamics appear to represent slab stag-
nations and their collapse processes (Fukao & Obayashi 2013, Fukao et al. 2009, Scire et al. 2017).
These images also suggest that there are four stages of the fate of subducted slabs: (a) stagna-
Mantle avalanche: tion above the 660-km discontinuity, (b) penetration of the 660-km discontinuity at slab depths of
large-scale cold 660–1,000 km, (c) trapping around a depth of 1,000 km, and (d) further penetration of the deeper
downwelling flow layers of the mantle (Fukao & Obayashi 2013). Many theoretical and numerical studies of the
across 660-km depth
caused by an interactions between mantle flow and the spinel–postspinel (ringwoodite–bridgmanite) transition
endothermic phase have provided interpretations of the various schematic images obtained by experimental mea-
transition and by surements at high-temperature and high-pressure conditions, as well as the imaging of the deep
Rayleigh-Taylor-type interior provided by global and regional seismic tomography, including (a) double-layered mantle
instability
convection (Christensen & Yuen 1984, Nakakuki et al. 1994), (b) mantle avalanches (Honda et al.
1993, Solheim & Peltier 1994, Tackley et al. 1993), and (c) more detailed physics of the formation

42 Nakagawa • Nakakuki
EA47CH03_Nakagawa ARjats.cls April 20, 2019 10:38

and collapse of slab stagnation across the 660-km depth (Nakakuki et al. 2010). According to more
realistic mineral physics information about the density crossover caused by the different phase
transition pressures between the spinel–postspinel transition and the garnet–postgarnet transition
Basalt barrier: barrier
(Irifune & Ringwood 1993, Ono et al. 2001), the mantle dynamics at depths of 660 to 720 km may in which, owing to
induce the basalt layering (basalt barrier) above the 660-km depth (Ballmer et al. 2015, Tackley density crossover
et al. 2005) and facilitate the generation and supply of a depleted component to the mantle transi- between olivine and
tion zone (Nakagawa & Buffett 2005). Realistic phase transition data about mantle minerals com- pyroxene phases,
basaltic material in
puted from the Gibbs free energy minimization have confirmed the presence of the basalt barrier
upwelling plumes
using various compositional models of mantle rocks (Nakagawa et al. 2010) and a detailed process could be segregated
for generating large-scale heterogeneity in whole mantle (Stixrude & Lithgow-Bertelloni 2012).
Dense hydrous
Moreover, in the mantle transition zone to the uppermost lower mantle, the water transported
magnesium silicate
via plate subduction can be absorbed (Iwamori 2007, Kawakatsu & Watada 2007). The water (DHMS): hydrous
transported into the mantle transition zone may be filtered by dense silicate melt at a depth of ap-
Annu. Rev. Earth Planet. Sci. 2019.47:41-66. Downloaded from www.annualreviews.org

mantle mineral
proximately 400 km in a process referred to as the transition zone water filter (Bercovici & Karato expressed by the
2003, Karato et al. 2006). This hypothesis may be confirmed with high-pressure mineral physics to H2 O-MgO-SiO2
Access provided by 186.183.193.242 on 05/03/20. For personal use only.

system, which is stable


discover and analyze diamond inclusions with high water content (Pearson et al. 2014) and inter-
from 10 to 80 GPa
pretations of seismological and electromagnetic images (Houser 2016, Kuvshinov 2012, Schmandt
et al. 2014), but it has not been well resolved by the numerical modeling of mantle dynamics with
water circulation. Based on petrological measurements (Iwamori 2004, 2007; Komabayashi et al.
2004) and the discovery of dense hydrous magnesium silicate (DHMS), which seems to be stable
under lower mantle convection conditions (Nishi et al. 2014, Ohira et al. 2014) and plays a sig-
nificant role in the hydrogen cycle in the deep lower mantle (Hu et al. 2016, Nishi et al. 2017),
the mantle transition zone and uppermost lower mantle can be interpreted as transporting large
amounts of water from cold and hydrous ocean plates. Recent progress made in numerical mod-
eling, including petrological measurements of the water solubility limits of each mantle mineral,
can be used to address water migration in the mantle and to understand the ways in which the wa-
ter transport efficiency of plate subduction is key for understanding the deep mantle water cycle
(Iwamori & Nakakuki 2013, Nakagawa & Iwamori 2017, Nakagawa et al. 2015, Nakao et al. 2016,
Richard et al. 2002).
Here, we provide a review of the current status of the numerical modeling of dynamics in
the uppermost lower mantle and the mantle transition zone across the 660-km depth, which has
mainly focused on the numerical modeling of the water cycle caused by subducting slabs (stagna-
tion and penetration) and global-scale mantle dynamics. This is because several general reviews
have recently addressed the dynamics of subducting slabs (e.g., Faccenda & Dal Zilio 2017). Again,
we do not provide a general review of the deep mantle water cycle based on mineral physics and
geophysical observations.

2. BASIC KNOWLEDGE
In this section, we describe some basic information that is important for understanding the dy-
namics of the mantle transition zone and the uppermost lower mantle in numerical modeling
approaches.

2.1. Mantle Layering


As indicated by various investigations of the numerical modeling of mantle convection, the
style of layering of mantle convection occurring at the 660-km discontinuity is related to
phase transitions and chemical structures that are mainly caused by the spinel–postspinel phase

www.annualreviews.org • Dynamics in the Uppermost Lower Mantle 43


EA47CH03_Nakagawa ARjats.cls April 20, 2019 10:38

transition (ringwoodite–bridgmanite) occurring at a depth of approximately 660 km (Christensen


& Yuen 1984, Nakakuki et al. 1994). This phase transition generally has a negative Clapeyron
slope (endothermic); thus, the temperature decreases at the phase boundary. This means that
the phase boundary shifts to a higher pressure, such that the convective flow may be decreased
across the phase boundary (e.g., Tackley 1995). However, the value of the Clapeyron slope of
this phase transition may have a great uncertainty, ranging from −3 to −0.5 MPa/K, as indicated
by recent mineral physics measurements (Fei et al. 2004, Hirose 2002, Katsura et al. 2003). For
smaller negative Clapeyron slopes, it may be difficult to generate the mantle layering with an
endothermic phase transition at the bottom of the mantle transition zone (e.g., Nakakuki et al.
1994). Additionally, as mentioned below, the increase in viscosity at the bottom of the mantle
transition zone may strongly affect the dynamics of the subducting lithosphere and global-scale
mantle dynamics (e.g., Tackley 1996).
The mantle layering across the 660-km discontinuity is strongly affected by the stagnation and
Annu. Rev. Earth Planet. Sci. 2019.47:41-66. Downloaded from www.annualreviews.org

penetration of the subducting slab. Numerical modeling approaches attempt to render the style of
layering in mantle dynamics (i.e., whole-mantle convection, two-layered convection, or another
Access provided by 186.183.193.242 on 05/03/20. For personal use only.

style) to be consistent with the various implications of the high-pressure and high-temperature
measurements of mantle minerals and the deep Earth imaging provided by geophysical (both
seismological and electromagnetic) observations (e.g., Fukao et al. 2009). However, with thermo-
chemical heterogeneities, the style of mantle convection may not be clearly differentiated between
two-layered and whole-mantle convection models. Such models can explain both the geochem-
ical and geophysical interpretations of deep mantle dynamics (Ballmer et al. 2017, Stixrude &
Lithgow-Bertelloni 2012, Tackley 2000).

2.2. Water Dependence on the Physical Properties of Mantle Material


As mentioned in Section 1, some petrological and mineral physics data have revealed the amount
of water that mantle rocks can absorb (e.g., Iwamori 2007). In addition, a few schematic images
have provided information about the material properties measured at high-pressure and high-
temperature conditions (e.g., Komabayashi et al. 2004). To reveal the physics behind the schematic
images obtained from experimental and observational investigations, it is important to determine
the rheological properties and water solubilities of hydrous mantle minerals.
First, the rheological properties of mantle rocks are measured in deformation experiments of
mantle minerals at high-pressure and high-temperature conditions (Karato & Wu 1993, Li et al.
2008); the viscosity of hydrous mantle minerals is usually given as
 
H (p)
η(T , p, fw ) = A fw−r exp , 1.
RT

where η is the viscosity of mantle material as a function of temperature, pressure, and water fu-
gacity; r is the exponential constant for the strength of the water dependence of viscosity; A is the
prefactor, which is determined based on a reference state; H = E + pV is the activation enthalpy;
R is the gas constant; and T is the mantle temperature. The water fugacity is empirically scaled
with the mantle water content, which is given by Li et al. (2008):

ln ( fw ) = a0 + a1 lnCOH + a2 ln2COH + a3 ln3COH , 2.

where a0 –a3 are the fitting constants; and COH (H/106 Si) is the amount of hydroxide component
in hydrous minerals, which is scaled with the mantle water content. The mantle water constant is

44 Nakagawa • Nakakuki
EA47CH03_Nakagawa ARjats.cls April 20, 2019 10:38

given by Komacek & Abbot (2016):

2 × 106Cw μ̃
COH = , 3.
1 − μ̃

where Cw is the mantle water content, and μ̃ is the ratio of the molar mass between olivine and
water. However, this scaling relationship has only been assessed in shallower regions by mineral
physics measurements; it has not been confirmed in deep mantle minerals. To avoid such uncer-
tainty, it is generally assumed that the viscosity of the hydrous mantle is defined as
 −r  
Cw Ew + pVw
ηw = Aw exp , 4.
Cw,ref RT

where Aw is the prefactor of the hydrous mantle viscosity; Cw,ref is the reference mantle water
Annu. Rev. Earth Planet. Sci. 2019.47:41-66. Downloaded from www.annualreviews.org

content (assumed to range from 110 to 620 ppm) (Arcey et al. 2005, Nakao et al. 2016); and Ew
and Vw are the activation energy and volume, respectively. Hydrous mantle minerals are generally
Access provided by 186.183.193.242 on 05/03/20. For personal use only.

suggested to be much weaker than dry mantle minerals, which is related to the exponential index r
shown in Equations 1 and 4. This value appears to range from 1.0 to 1.2 (e.g., Hirth & Kohlstedt
1996, Mei & Kohlstedt 2000) or, sometimes, to be much higher (∼2.0) (Korenaga & Karato 2008).
However, recent experiments on hydrogen diffusion in mantle minerals (Fei et al. 2013) have
indicated that the water dependence of viscosity may not be very sensitive to the mantle water
content because of the small values of the exponent associated with the water content effects (see
Equations 1 and 4), which are ∼0.3. The typical rheological properties of dry and hydrous mantle
minerals are listed in Table 1.
Second, the numerical modeling of mantle dynamics should incorporate the water solubility of
hydrous mantle minerals. Based on thermodynamic measurements of the maximum water content
of hydrous mantle minerals, Figure 1a shows water solubility maps of ambient mantle rocks, such
as that of mantle peridotite given by Iwamori (2007) and those compiled by Komabayashi et al.
(2004), which are used to assess the detailed solubility conditions of upper mantle minerals. The
results of the numerical modeling of the mantle water cycle with whole-mantle convection per-
formed by Nakagawa et al. (2015), Nakagawa & Spiegelman (2017), and Nakagawa & Iwamori
(2017) are included as a way to assess the upper mantle solubility maps provided by petrologi-
cal and mineral physics data, and the lower mantle solubility is assumed to be a fixed value of
100 ppm (Karato 2011). Note that the water solubility limit of ambient lower mantle minerals,
such as bridgmanite, may have a large degree of uncertainty, ranging from 10 ppm to 0.2 wt%
(Bolfan-Casanova 2005, Karato 2011, Murakami et al. 2002, Panero et al. 2015). In addition, when

Table 1 Typical rheological parameters of hydrous mantle minerals


Parameter Definition Range
Ed,diff Activation energy for diffusion creep of dry olivine 250–300 kJ/mol
Ed,disl Activation energy for dislocation creep of dry olivine 540–610 kJ/mol
Ew,diff Activation energy for diffusion creep of wet olivine 240–370 kJ/mol
Ew,disl Activation energy for dislocation creep of wet olivine 430–580 kJ/mol
r Exponential index 0.3–1.98

Data on activation energy taken from Karato & Wu (1993) and Korenaga & Karato (2008); data on the exponential index
of the water dependence of viscosity taken from Fei et al. (2013), Mei & Kohlstedt (2000), and Korenaga & Karato (2008).
Since the activation volumes of the deformation of olivine are quite difficult to determine using experimental data, they are
not listed in this table.

www.annualreviews.org • Dynamics in the Uppermost Lower Mantle 45


EA47CH03_Nakagawa ARjats.cls April 20, 2019 10:38

Without DHMS With DHMS


4,000

3,500
a b
3,000 therm
Temperature (K)

e geo
Plum
2,500
ged
2,000 Avera

1,500 eotherm
Slab g
1,000

500
0 20 40 60 80 100 120 0 20 40 60 80 100 120
Pressure (GPa) Pressure (GPa)
Annu. Rev. Earth Planet. Sci. 2019.47:41-66. Downloaded from www.annualreviews.org

log10 water solubility (wt%)

–3 –2 –1
Access provided by 186.183.193.242 on 05/03/20. For personal use only.

0 1 2

Figure 1
Water solubility maps (a) of ambient mantle rocks and (b) expanded to include the temperature and pressure range of the entire mantle.
Figure adapted from Nakagawa et al. (2018). Abbreviation: DHMS, dense hydrous magnesium silicate.

expanding to whole-mantle conditions, the DHMS solubility, which is stable at uppermost lower
mantle conditions, must be added to the upper mantle solubility map (Nakagawa et al. 2018). The
maximum solubility of DHMS can be estimated based on its chemical formula and is ∼12 wt%
for phase H (Nishi et al. 2014, Ohira et al. 2014). Incorporating the DHMS found in the upper-
most mantle into the upper lower mantle, Figure 1b shows the water solubility map expanded
to the temperature and pressure range of the entire mantle. The uppermost lower mantle and
mantle transition zone can be interpreted as large reservoirs of water transported via plate sub-
duction and are crucial for understanding the mantle water cycle using data from mineral physics
measurements and geophysical observations.
Most of the numerical and theoretical models of mantle water circulation do not account for the
generation of excess water seen on water solubility maps, which can be described by the regassing–
degassing relationship (e.g., Crowley et al. 2011) or using a very simplified profile of water solu-
bility (e.g., Richard et al. 2002). However, recent progress has been made in modeling the mantle
water cycle at the scales of both slab subduction and global mantle dynamics (Iwamori & Nakakuki
2013, Nakagawa et al. 2015, Nakao et al. 2016), and in successfully incorporating the generation
of excess water in the deep mantle into the numerical modeling of mantle convection with the
tracer particle approach. Figure 2 shows a schematic illustration of the excess water migration
taken from Nakagawa et al. (2015). When the mantle water content exceeds the water solubility
of mantle minerals, excess water can be generated and vertically transported above each grid point
until the excess water can no longer be generated during one time step. Alternatively, the two-
phase flow approach can also include such physics with a more realistic procedure (Horiuchi &
Iwamori 2016, Wilson et al. 2014). However, with current computational power, such an approach
is very time-consuming for global-scale numerical modeling with high resolution (approximately
1–10-km resolution near the surface) because the fluid migration dynamics using the tracer parti-
cle approach may use a similar timescale as the two-phase flow model (Nakagawa & Spiegelman
2017). This means that the completed two-phase flow model is applicable to the wedge mantle
dynamics caused by slab subduction, whereas a conventional approach can still be used for global-
scale mantle water circulation.

46 Nakagawa • Nakakuki
EA47CH03_Nakagawa ARjats.cls April 20, 2019 10:38

Water content assigned


in tracer particles

Mapping to numerical grids

To the surface To the surface

Excess Excess
Max Max
water water

TIME TIME
Excess Excess
Max Max Max Max
water water
Annu. Rev. Earth Planet. Sci. 2019.47:41-66. Downloaded from www.annualreviews.org

Excess Max Max


water
Access provided by 186.183.193.242 on 05/03/20. For personal use only.

Numerical grids to each tracer

Water content assigned


in tracer particles

Figure 2
Schematic illustration of excess water migration using the tracer particle approach. When the mantle water
content exceeds the water solubility of mantle minerals, the excess water can be generated and vertically
transported above each grid point until the excess water can no longer be generated during one time step.
Figure adapted from Nakagawa et al. (2015).

In addition to these requirements, there are two other important properties of hydrous mantle
minerals, i.e., the effects of density and solidus temperatures. First, hydrous mantle minerals are
generally less dense than dry mantle minerals by 0.1% to 1% (e.g., Wang et al. 2006). A recent
estimate indicated that the mineral density can decrease by ∼1.4 wt% with an additional 1 wt% of
mantle water content (e.g., Ye et al. 2012). In numerical modeling studies concerning the density
reduction caused by hydrous mantle minerals, the density of hydrous mantle is given as

ρ = ρ0 [α(T − T0 ) − βCw ] , 5.

where ρ0 is the reference density; α is the thermal expansivity; T0 is the reference temperature;
Cw is the mantle water content (in weight percent); and β is the density reduction rate due to the
mantle water content, ranging from 1.2 to 1.6, but this may be strongly uncertain (Nakao et al.
2016, Richard & Iwamori 2010).
Second, hydrous mantle minerals have lower solidus temperatures than dry mantle minerals
(e.g., Kawamoto & Holloway 1997). Compiling various experimental data, Katz et al. (2003) cre-
ated a scaling law of solidus temperatures based on the mantle water content, which is given as

Tw = 43(Xw )0.75 , 6.

www.annualreviews.org • Dynamics in the Uppermost Lower Mantle 47


EA47CH03_Nakagawa ARjats.cls April 20, 2019 10:38

4,500 10 2
a b
Solidus temperature (K)

4,000
10 1

Maximum water
3,500

content (wt%)
3,000
10 0
2,500

2,000
Dry 10 –1
1,500 Wet

1,000 10 –2
0 500 1,000 1,500 2,000 2,500 0 500 1,000 1,500 2,000 2,500
Depth (km) Depth (km)
Figure 3
Annu. Rev. Earth Planet. Sci. 2019.47:41-66. Downloaded from www.annualreviews.org

(a) Solidus temperature profile with the influence of mantle water content. (b) Maximum mantle water content used for the solidus
temperature shown in panel a, which includes the dense hydrous magnesium silicate solubility taken from Nakagawa et al. (2018).
Access provided by 186.183.193.242 on 05/03/20. For personal use only.

where Xw is the mantle water content (in weight percent). Figure 3 shows the sensitivity of the
solidus temperature to the mantle water content. Using the maximum water content expected
from hydrous mantle convection simulations (e.g., Nakagawa et al. 2018), the solidus temper-
ature can be reduced to ∼100 K. However, such a scaling law is valid only at upper mantle
pressure conditions (up to 10 GPa); thus, this scaling relationship is not widely applicable to the
pressure range of the entire mantle. It should be revised so that the scaling law of the solidus
temperature as a function of the mantle water content is applicable under the temperature and
pressure range of the entire mantle.

3. WATER IN THE MANTLE TRANSITION ZONE AND UPPERMOST


LOWER MANTLE: ROLE OF THE STAGNANT SLAB AND ITS
PENETRATION
In this section, we review the numerical modeling of slab dynamics in mantle convection simu-
lations, including (a) the formation and collapse mechanisms of slab stagnation, along with the
significant phenomena of mantle dynamics in the mantle transition zone and uppermost lower
mantle, and (b) hydration in the mantle transition zone and the transportation of water in the
deep mantle associated with the collapse of stab stagnation.

3.1. Formation and Collapse Mechanisms of Slab Stagnation


In the four stages of the fate of subducting slabs mentioned in Section 1, two mechanisms are
required to generate slab stagnation and penetration at the bottom of the mantle transition zone:
Trench rollback:
tectonic phenomenon (a) the viscosity jump at a depth of 660 km and (b) the trench rollback effect (e.g., Goes et al. 2017).
in which an oceanic Since the Clapeyron slope of the spinel–postspinel phase transition may have great uncertainty
trench moves in the ranging from −3.0 to −0.5 MPa/K (Fei et al. 2004, Hirose 2002, Houser & Williams 2010, Ito &
direction opposite to Takahashi 1989, Katsura et al. 2003), it is necessary to incorporate other physics into the modeling
plate motion caused by
of mantle dynamics to assess the formation mechanism of slab stagnation suggested by global to-
the subducting slab
mographic images. Again, we focus on recent progress made in numerical modeling of the mantle
dynamics of the uppermost lower mantle and the mantle transition zone.

48 Nakagawa • Nakakuki
EA47CH03_Nakagawa ARjats.cls April 20, 2019 10:38

In their attempts to understand the formation of slab stagnation in numerical modeling ap-
proaches, previous studies (e.g., Tagawa et al. 2007) indicated that trench rollback seemed to be
essential in generating the slab stagnation at the bottom of the mantle transition zone in numerical
Metastable olivine:
models of mantle convection, but that a negative Clapeyron slope was still required to achieve slab temporarily stable
stagnation. To assess the effects of these two processes on slab stagnation, Torii & Yoshioka (2007) olivine in temperature
conducted simple numerical modeling of subducting slabs in a mantle convection system with a and pressure
wide range of parameters to investigate the sensitivity of slab stagnation associated with the ve- conditions in which
high-pressure minerals
locity of trench rollback and the viscosity jump, as well as the Clapeyron slope of the bridgmanite
such as wadsleyite or
transition, which indicated the possible range of slab stagnation with the small Clapeyron slope ringwoodite are stable
of the spinel–postspinel phase transition. Recent investigations of the interaction between sub-
ducting slabs and the mantle transition zone with more realistic material properties have revised
our understanding of the mechanism of slab stagnation at the bottom of the mantle transition
zone (Agrusta et al. 2017, Cizkova & Bina 2013). In particular, trench rollback, along with the
Annu. Rev. Earth Planet. Sci. 2019.47:41-66. Downloaded from www.annualreviews.org

viscosity jump at a depth of 660 km, seems to play a significant role in generating slab stagnation
(Cizkova & Bina 2013, Goes et al. 2017). Moreover, the age of the oceanic lithosphere is also a
Access provided by 186.183.193.242 on 05/03/20. For personal use only.

key component of the generation of slab stagnation (Agrusta et al. 2017).


The fate of slab stagnation in a mantle convection system, which is the collapse of a stagnant
slab (or slab penetration), can be interpreted as a mantle avalanche event in global-scale mantle
convection simulations with multiple phase transitions (Honda et al. 1993, Tackley et al. 1993).
With recent progress in numerical modeling techniques and computational power, more detailed
information about the processes of the collapse events of stagnant slabs is available. Earlier studies
(Goes et al. 2008, Yoshioka & Naganoda 2010) demonstrated slab penetration into the lower
mantle and pointed out that the low-viscosity zone below the mantle transition zone seems to
play an important role in both slab stagnation and penetration. Honda (2016) tried to explain the
seismic image indicating slab penetration under northeast China and pointed out that the age
of the oceanic lithosphere and plate velocity exert stronger controls on slab penetration events
rather than negative Clapeyron slopes. However, the origin of the 660-km discontinuity is still
quite controversial, and it is unclear whether this discontinuity is caused by a phase transition or
by chemical layering (Arredondo & Billen 2016, Nakakuki et al. 1994). More careful investigations
of this topic are expected in future studies.
For more physical interpretations of slab stagnation and penetration (collapse), it is essential to
use the force balance associated with a buoyancy at the bottom of the mantle transition zone (Billen
2010, Nakakuki et al. 2010). Figure 4 schematically illustrates such a force balance in a subducting
slab; this force balance is essential for understanding the physics of stagnation and collapse process.
Figure 5 shows the evolution of the stress field during the process leading from the stagnation
induced by trench migration to the penetration caused by gravitational instability (Nakakuki et al.
2010). The slab rollback causes concentration of the slab deformation in the middle section and
reduction of the stress in the deep section of the slab (Figure 5a). The stress at the hinge of the
slab is increased by a viscous resistance of the ambient mantle to the horizontally lying section
of the slab (Figure 5b,c). Growth of the slab portion that protrudes into the lower mantle finally
generates the slab penetration (Figure 5d). This sequence of slab stress evolution corresponds to
the progressive depth variation of deep-focus earthquakes (Fukao & Obayashi 2013).
The buoyancy force in a subducting slab may be strongly influenced by the following two
material properties, in addition to temperature and chemical effects: (a) the metastability at the
olivine–wadsleyite phase transition (Rubie 1984) and (b) the grain-size reduction accompanying
the olivine–wadsleyite and ringwoodite–bridgmanite phase transitions (Karato et al. 2001).
First, the metastable olivine produced by the suspension of the olivine–wadsleyite phase transi-
tion generates a less dense wedge in the slab (Schmeling et al. 1999). These effects are expected to

www.annualreviews.org • Dynamics in the Uppermost Lower Mantle 49


EA47CH03_Nakagawa ARjats.cls April 20, 2019 10:38

Vt fw
Vs
fp
fs
fr fz

fv fH
fn
ls
Annu. Rev. Earth Planet. Sci. 2019.47:41-66. Downloaded from www.annualreviews.org

Figure 4
Schematic illustration of the force balance of a slab at a 660-km depth. Important quantities determining the
stagnation or penetration of the slab are the upward buoyancy force by the phase transition ( fp ); the
Access provided by 186.183.193.242 on 05/03/20. For personal use only.

downward buoyancy force generated by the horizontal section of the slab ( fn ); and the downward
component ( fz ) of the viscous force ( fs ) generated by the inclined section of the slab, which may cause
Rayleigh-Taylor-type instability. A growth timescale of this type of instability can be written as τRT ηeff /σeff ,
where ηeff is the effective slab viscosity and σeff = ( fz + fn − fp )/ls , with ls the length of the stagnated slab
at 660 km. fH and fV show the horizontal and vertical viscous resistance from the ambient mantle,
respectively, and Vs and Vt show the velocities of subducting plate motion and trench migration, respectively.
Figure adapted with permission from Nakakuki et al. (2010).

enhance slab stagnation by reducing the negative buoyancy of the slab. Additionally, the reduction
of the negative buoyancy slows the descent of the slab. Decreasing the speed of slab descent tends
to promote slab rollback, which is the most efficient mechanism for generating a stagnant slab.
Kubo et al. (2009) confirmed that metastable olivine may strongly affect slab stagnation under the
small negative value of the Clapeyron slope found in the uppermost lower mantle in the Mari-
ana slab using the numerical modeling of slab dynamics and the incorporation of disequilibrium
phase changes such as the metastable effect. In addition, Agrusta et al. (2014) also demonstrated
that slab stagnation occurs with metastable pyroxene. This style of stagnation appears to not be
very different from that found in the metastable olivine phase system (ringwoodite). Therefore,
the disequilibrium phase transition may be quite significant for revealing slab stagnation with a
small negative Clapeyron slope because this metastability may reduce the buoyancy force asso-
ciated with the subducting slab. Second, the grain-size reduction effect may decrease the viscos-
ity of the slab center. Karato et al. (2001) proposed that slab softening due to grain-size reduc-
tion is a key mechanism of slab stagnation and penetration. However, numerical modeling has
indicated that the grain-size reduction does not significantly affect slab stagnation because the
grain growth is diminished only in the cold core of the slab (Cizkova et al. 2002, Nakakuki et al.
2010, Tagawa et al. 2007). In addition, Figure 5 shows that slab penetration may occur with a
grain-size reduction caused, under high-stress conditions, by slab protrusion under the 660-km
discontinuity.
In more realistic mantle material models, mantle rock can be decomposed into two
phase transition systems, i.e., olivine–spinel–bridgmanite–postperovskite and pyroxene–garnet–
bridgmanite–postperovskite (e.g., Irifune & Ringwood 1993, Murakami et al. 2004, Oganov &
Ono 2004). Between these two phase transition systems, the transition pressure of bridgmanite is
different, which results in a density crossover (see Figure 6a). By incorporating such effects, the

50 Nakagawa • Nakakuki
EA47CH03_Nakagawa ARjats.cls April 20, 2019 10:38

22.3 Myr 48.7 Myr


(cm/year) 15
Surface
velocity
10 a c
5
0
−5
0 log10 viscosity
structure (°C)

26
Viscosity

500 24
22
1,000 20
18
1,500 16
0 Maximum
stress (MPa)

stress (MPa)
Maximum

500 200

100
Annu. Rev. Earth Planet. Sci. 2019.47:41-66. Downloaded from www.annualreviews.org

1,000

1,500 0
Access provided by 186.183.193.242 on 05/03/20. For personal use only.

29.4 Myr 60.6 Myr


15
(cm/year)
Surface
velocity

10 b d
5
0
−5
0 log10 viscosity
structure (°C)

26
Viscosity

500 24
22
1,000 20
18
1,500 16
0 Maximum
stress (MPa)

stress (MPa)
Maximum

500 200

1,000 100

1,500 0
4,000 5,000 6,000 7,000 8,000 4,000 5,000 6,000 7,000 8,000
Horizontal direction (km) Horizontal direction (km)
Figure 5
Evolution of a subducted slab with weakening due to grain-size reduction at the 660-km phase transition at (a) 22.3, (b) 29.4, (c) 48.7,
and (d) 60.6 Myr, showing the (top graphs) surface velocity; (middle graphs) viscosity structure (color gradient) with the dashed horizontal
lines at 1,000, 1,200, and 1,400°C; and (bottom graphs) maximum stress. Horizonal dashed lines in viscosity and mantle water profiles are
positions of phase transition on olivine-spinel and spinel-bridgmanite. Figure adapted with permission from Nakakuki et al. (2010).

basalt barrier above a depth of 660 km can be found as a result of the density crossover between
the spinel–postspinel and garnet–bridgmanite transitions using simple material models (Tackley
et al. 2005) and thermodynamic-based material models (Nakagawa et al. 2010). An example of a
basalt barrier caused by a density crossover is shown in Figure 6b. This basalt barrier seems to
have some impact on the circulation of material across the 660-km depth that is caused by the
counterflow of slab penetration (Nakagawa & Buffett 2005). Ballmer et al. (2015) pointed out that
a basalt barrier may be interpreted as explaining the origin of the seismic anomalies observed be-
neath the Hawaiian plumes. Kaneshima (2009) and Waszek et al. (2018) also pointed out that the
scattered features found in seismic observations of the lower mantle correspond to the distribution
of oceanic crust caused by slab penetrations.

www.annualreviews.org • Dynamics in the Uppermost Lower Mantle 51


EA47CH03_Nakagawa ARjats.cls April 20, 2019 10:38

0
Ba Basalt
200
Ha
rz sa Density a b
Depth (km)

bu lt barrier
rg crossover
400 i te Harzburgite
600 2

800
1,000
3,000 3,500 4,000 4,500 5,000 5,500
Density (kg/m3)
0
1
200 Basalt
Depth (km)

400
600 Basalt
Basalt barrier
800
Annu. Rev. Earth Planet. Sci. 2019.47:41-66. Downloaded from www.annualreviews.org

1,000
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9
Access provided by 186.183.193.242 on 05/03/20. For personal use only.

Basalt composition
Figure 6
(a) Reference density profiles of olivine- and pyroxene-based phase transitions. (b) Chemical structure computed from thermochemical
mantle convection (Ballmer et al. 2015). This numerical modeling assumes a dry mantle.

3.2. Formation of Hydrous Mantle Transition Zone Caused by Slab Stagnation


Cold subducting slabs play a significant role in transporting water into the deep mantle; how-
ever, it is still unclear how hydrous minerals affect the dynamics of subducting slabs in the mantle
transition zone and uppermost lower mantle (Iwamori 2007, Kawakatsu & Watada 2007). From
an observational point of view, the water content in the mantle transition zone can be computed
based on the seismic anomalies of a global tomography model; this content is ∼1.0 wt% (Suetsugu
et al. 2006). However, a physical mechanism of generating such a hydrated region has not yet
been discussed. Ye et al. (2011) found a water-bearing region caused by slab stagnation in seis-
mic images of the big mantle wedge beneath the East Asia region (Chen et al. 2017, Ye et al.
2011), which was also indicated by electrical conductivity modeling (Ichiki et al. 2006, Kuvshinov
2012). The formation mechanism of the big mantle wedge has been reconciled with the numeri-
cal modeling of slab subductions performed by Kameyama & Nishioka (2012) and Honda (2016,
2017), which suggested the occurrence of trench rollback effects. However, their studies have
not discussed the hydrous region on the edge of the big mantle wedge. To explain the hydration
mechanism in the big mantle wedge, Richard & Iwamori (2010) and Nakao et al. (2016) used a
dehydration reaction from a stagnant slab that may be sufficiently cold that a hydrous slab with
large water content can pass through the choke point (Iwamori 2004, Komabayashi et al. 2004).
From a mineral physics point of view, hydrous mantle minerals are generally less dense than dry
mantle minerals (Angel et al. 2001, Jacobsen et al. 2008); this factor also seems to play an im-
Choke point: point of
maximum water portant role in slab stagnation (Richard & Bercovici 2009, Richard & Iwamori 2010). The main
solubility in which the role of mantle minerals that may host the water is the wadsleyite found in the mantle transi-
cold slab, including tion zone (e.g., Smyth & Kawamoto 1997). As wadsleyite’s role is well understood in the com-
hydrous oceanic crust, munity, we do not discuss this mineral further; however, it is important to note that wadsleyite
has to release water
plays an important role in absorbing the water in the mantle transition zone. Other numerical
investigations have also reproduced the hydrous region associated with the big mantle wedge,
which may indicate that the dehydration reaction occurred in the bottom of the mantle transition
zone (Faccenda 2014, He 2017, Wang et al. 2018). Figure 7 shows an example of the numerical

52 Nakagawa • Nakakuki
EA47CH03_Nakagawa ARjats.cls April 20, 2019 10:38

16.67 Myr
0
log10 viscosity
(Pa·s)
200 25

24
400
23

600 22

21
800
20

1,000 19

18
a
Depth (km)

1,200
Annu. Rev. Earth Planet. Sci. 2019.47:41-66. Downloaded from www.annualreviews.org

17
Water content
Access provided by 186.183.193.242 on 05/03/20. For personal use only.

(kg/kg)
200 0.06

400 0.05

0.04
600

0.03
800
0.02
1,000
0.01
1,200
b
0
5,000 5,200 5,400 5,600 5,800 6,000 6,200 6,400 6,600 6,800 7,000 7,200 7,400
Horizontal position (km)
Figure 7
(a) Viscosity and (b) mantle water profiles computed from the numerical modeling of slab subduction. Dashed lines in both the viscosity
and water profiles correspond to the positions of olivine-spinel and spinel-bridgmanite transitions. Figure adapted from Nakao et al.
(2016).

results of the slab dynamics obtained from hydrous mantle convection simulations (Nakao et al.
2016). Nakao et al.’s (2016) investigation demonstrated the importance of the rheological proper-
ties and density reduction due to the mantle water content in generating slab stagnation and trench
rollback.
After the stagnant slab collapses in a given region, some amount of water can still be trans-
ported into the deep mantle because the slab geotherm seems to be sufficiently low that the
DHMS can absorb the water from the collapsed slab (e.g., Nishi et al. 2014). Interpreting
the water in the deep mantle caused by slab stagnation and penetration, some mineral physics
measurements have implied a phase transition of hydrous iron oxides (Hu et al. 2016, Nishi et al.
2017), in addition to the mineral physics interpretations of water in the mantle transition zone
(Bercovici & Karato 2003, Karato et al. 2006). However, in numerical modeling, the implications
provided by mineral physics have not yet been investigated, and this should be done in future
studies. It is difficult to reproduce mineral physics interpretations in numerical models. For
instance, the transition zone water filter hypothesis requires partial melting to occur at a depth
of 400 km because the melted material at that pressure may have a greater density than the

www.annualreviews.org • Dynamics in the Uppermost Lower Mantle 53


EA47CH03_Nakagawa ARjats.cls April 20, 2019 10:38

surrounding solid mantle (see Bercovici & Karato 2003) because the handling of the density
crossover between solid and partial molten material at a depth of 400 km cannot be implemented
in a very straightforward manner; however, promisingly, some numerical code could potentially
address the complicated relationship between the densities of solid and partially molten material
(e.g., Dannberg & Heister 2016). Moreover, in modeling deep mantle hydration, the redox state
is key because hydrous iron oxide seems to be an essential host of water in the lower mantle (e.g.,
Rohrbach & Schmidt 2011) but is not currently included in numerical models. Modeling the
water solubility of lower mantle minerals, i.e., bridgmanite, is key for understanding the hydration
process in the mantle transition zone. Again, this may have large uncertainty, ranging from 10 ppm
to 2,000 ppm (Bolfan-Casanova 2005, Karato 2011, Murakami et al. 2004, Panero et al. 2015). To
constrain the water solubility of the ambient lower mantle, Nakagawa (2017) numerically mod-
eled hydrous mantle convection with variation in the water solubility of lower mantle material.
This modeling suggested that, to find the hydrated mantle transition zone, the water solubility
Annu. Rev. Earth Planet. Sci. 2019.47:41-66. Downloaded from www.annualreviews.org

of lower mantle material should preferably be a few hundreds of parts per million, rather than
∼1,000 ppm.
Access provided by 186.183.193.242 on 05/03/20. For personal use only.

For example, Figure 8 shows the time dependence of the one-dimensional horizontally av-
eraged mantle water content structure, including the effects of DHMS solubility. In the mantle
transition zone, the hydrated region appears in approximately 500 million years and spreads out
to the entire mantle transition zone within an additional 2 billion years, so that the water content
is consistent with that provided by geophysical observations. After the hydrated mantle transition
zone is formed, the upper lower mantle, including the DHMS, can be hydrated.
In addition to the hydration of the mantle transition zone caused by slab stagnation and pen-
etration, some numerical models of subduction with water migration have addressed the global-
scale water budget because it is essential to constrain the amount of water that can be transported
by cold subducted slabs to the deep mantle (e.g., Iwamori 2007, van Keken et al. 2011). However,

3.0

2.5 Mantle DHMS Time (Ga)


transition (uppermost and 0.5
zone upper lower mantle) 1.0
Water content (wt%)

2.0 2.0
3.0
4.0
1.5 4.6

1.0

0.5

0
0 20 40 60 80 100 120
Pressure (GPa)
Figure 8
Time dependence of one-dimensional horizontally averaged profiles of mantle water content for a case
including dense hydrous magnesium silicate (DHMS) solubility. With increasing time, the mantle transition
zone becomes more hydrated due to excess water migration.

54 Nakagawa • Nakakuki
EA47CH03_Nakagawa ARjats.cls April 20, 2019 10:38

there have been many attempts at numerical mantle convection simulations with water migration
(e.g., Fujita & Ogawa 2013; Nakagawa & Iwamori 2017; Nakagawa et al. 2015, 2018; Richard et al.
2002); more detailed descriptions of the global-scale water cycle in the modeling of global-scale
mantle dynamics are provided in the next section.

4. DEEP MANTLE WATER CYCLE: GLOBAL-SCALE MANTLE


DYNAMICS
4.1. Rheological Effects
In this section, we discuss the rheological effects caused by mantle water content and their influ-
ences on global-scale dynamics in the hydrous mantle based on insights provided by numerical
modeling. The effects of the rheological properties of hydrous mantle rocks have been discussed
Annu. Rev. Earth Planet. Sci. 2019.47:41-66. Downloaded from www.annualreviews.org

using a simple semi-theoretical model of plate motion induced by hydrous mantle convection
(Crowley et al. 2011). This model indicates that the rheological dependence of mantle rocks on
Access provided by 186.183.193.242 on 05/03/20. For personal use only.

temperature and water content can enhance the heat transfer of mantle convection or water ab-
sorption in the deep mantle. In more detail, this model assesses the time-derivative nature of
viscosity, given as

dη(T , Cw ) ∂η dT ∂η dCw
= + = ηT Ṫ + ηW Ċw , 7.
dt ∂T dt ∂Cw dt
where η(T , Cw ) is the temperature- and water-dependent viscosity. The strength factor, which
is the ratio of the variation in viscosity caused by the mantle water content to the temperature-
dependent viscosity, can be written as
 
 η Ċ 
 W w
SWT =  , 8.
 ηT Ṫ 

where there is competition between (a) the viscosity reduction due to the increase in the water
content and (b) the increase in viscosity due to the decrease in temperature during the secular evo-
lution of the mantle. When this factor is larger than unity, mantle convection occurs in such a way
that its strength increases with time because the reduction in viscosity caused by the water input
is greater than the increase caused by secular cooling, which further enhances water absorption in
the mantle. In contrast, mantle cooling is enhanced when this factor is less than unity. This factor
may be useful for assessing which physical processes (thermal feedback or water enhancement)
are more effective in hydrous mantle convection.
In the full model of the dynamics of hydrous mantle convection, including water solubility
maps, the heat transfer of hydrous mantle convection is more effective than that of dry mantle
convection, which is ∼30%, because the mantle water content may reduce the viscosity and en-
hance the convective strength associated with the viscosity reduction (e.g., Nakagawa & Iwamori
2017, Nakagawa et al. 2015). Confirming the physical mechanism of mantle cooling and water
absorption, Figure 9 shows the strength factor defined in Equation 8, i.e., the mass-averaged
mantle temperature and mantle water mass as a function of time, as computed from the numeri-
cal simulations of hydrous mantle convection, for the case used by Nakagawa et al. (2018), and
the water-dependent viscosity is applied to Equation 4. This indicates that, under most con-
ditions, this value is less than unity. This may suggest that the strength of mantle convection
hardly increases over time with increases in the water content, even though the water input can
reduce the viscosity of the mantle. Rather, the reduction in viscosity due to water input helps

www.annualreviews.org • Dynamics in the Uppermost Lower Mantle 55


EA47CH03_Nakagawa ARjats.cls April 20, 2019 10:38

102

101

Enhancing the water cycle


10 0
Enhancing the thermal feedback
Strength factor
10 –1

10 –2

10 –3

10 –4
Annu. Rev. Earth Planet. Sci. 2019.47:41-66. Downloaded from www.annualreviews.org
Access provided by 186.183.193.242 on 05/03/20. For personal use only.

0 0.5 1.0 1.5 2.0 2.5 3.0 3.5 4.0 4.5


Time (Ga)
Figure 9
Strength parameter computed from the fully dynamical model of hydrous mantle convection (Nakagawa
et al. 2018).

accelerate secular cooling by increasing the surface heat flow. The rapid increase in the mantle
water mass is caused by sufficiently cooling the mantle temperature so that the mantle transi-
tion zone or DHMS region may absorb significant amounts of the water transported via plate
subduction.
The rheological properties of hydrous mantle minerals are still controversial because two dif-
ferent sets of experimental data have indicated two different extreme end members of the viscosity
dependence of mantle water content, and this is associated with the uncertainty of the exponential
index r in Equation 1. These end members include (a) the weak dependence (r ∼ 0.3) of the mantle
water content suggested by the diffusion measurements of Fei et al. (2013) and (b) the extremely
strong dependence of the mantle water content (r ∼ 2.0) suggested by the statistical data analysis of
deformation experiments by Korenaga & Karato (2008). Assuming weak dependence of the man-
tle water content in numerical models, the main features of the thermochemical structures under
hydrous mantle convection conditions are very similar to those of dry ones (Nakagawa et al. 2015).
In contrast, very high viscosity can be expected in the lower mantle with strong water-dependent
viscosity, which is somewhat unrealistic compared to the viscosity measurements observed with
geoid and sea-level change (∼1024 Pa·s). The rheological dependence of mantle water content
may also affect the styles of slab dynamics across the 660-km depth (Nakao et al. 2016), indicating
that slab stagnation is more likely to occur with the stronger rheological dependence of mantle
water content than with a weaker dependence. Moreover, in the lower mantle, the rheological
dependence of the mantle water content is not yet well understood because of the difficulties in
performing high-pressure and high-temperature experiments on material deformation. However,
recent experimental progress can address the deformation properties of silicates under the upper-
most lower mantle conditions (Girard et al. 2016). When the rheological properties of hydrous
minerals at lower mantle conditions are better clarified, we will be able to better understand the
rheological feedback of hydrous mantle convection by incorporating data from mineral physics
into numerical models.

56 Nakagawa • Nakakuki
EA47CH03_Nakagawa ARjats.cls April 20, 2019 10:38

4.2. Deep Mantle Water Cycle: Effects of Hydrous Phases on the Uppermost
Lower Mantle
In Section 4.1, we discuss the mantle water cycle with models of full mantle dynamics that incorpo-
rate realistic water solubility limits and rheological properties. Information about the deep mantle
water cycle can be inferred from water input by plate subduction in the mantle wedge (Iwamori
2007, Iwamori & Nakakuki 2013, Nakao et al. 2016, Richard & Iwamori 2010, van Keken et al.
2011). The above-cited investigations pointed out that constraining water storage in each hydrous
mantle mineral was important to explain the formation mechanisms and seismic images beneath
island arcs and inferred that the global water cycle occurred in the deep mantle but was not fully
understood. Several investigations have tried to reveal the global water cycle in the deep mantle
and its influence on the evolution of surface seawater, that is, the survival time of surface seawater
(Crowley et al. 2011, Korenaga 2011, Rüpke et al. 2004, Sandu et al. 2011); however, these studies
Annu. Rev. Earth Planet. Sci. 2019.47:41-66. Downloaded from www.annualreviews.org

did not include the water solubility effects of hydrous mantle minerals. To model the global-scale
mantle water cycle, Richard et al. (2002) initiated a global-scale mantle convection simulation with
Access provided by 186.183.193.242 on 05/03/20. For personal use only.

water migration based on a simplified treatment of the water solubility of hydrous mantle minerals,
which indicated that the mantle transition zone may hold a very large amount of water. Nakagawa
et al. (2015) fully investigated the effects of the water solubility of hydrous mantle minerals as
functions of temperature and pressure in a numerical mantle convection model by incorporating
the various complicated effects occurring in a convecting mantle. In addition, the water solubility
map of hydrous mantle minerals also placed constraints on the water absorbed in the early Earth’s
mantle (Nakagawa & Spiegelman 2017) and the evolution of surface seawater caused by the burst
of mantle water content, as shown in the detailed water solubility map of upper mantle minerals
(Nakagawa & Iwamori 2017). Despite the great progress made in modeling the global-scale water
cycle in mantle dynamics, the water solubilities of lower-mantle minerals have not been well con-
strained by mineral physics (Karato 2011, Panero et al. 2015). It is important to constrain the water
solubility in the lower mantle to reveal the water content in the mantle transition zone, which is
required to have a smaller solubility to achieve a consistent amount of water (Nakagawa 2017).
To resolve this issue, the recent discovery of DHMS (e.g., Phase H; Nishi et al. 2014), which is
stable under uppermost lower mantle conditions, is a key finding but has not been examined using
numerical models of hydrous mantle convection. Nakagawa et al. (2018) attempted to investigate
the influences of DHMS on the evolution of both surface seawater and mantle water content.
Figure 10 shows the mantle water content both with and without DHMS and the time series of
surface seawater and the mantle water mass taken from Nakagawa et al. (2018). With DHMS, the
high water content region may be extended to the uppermost lower mantle, corresponding to a
cold slab penetrating the lower mantle, but this does not occur in the case without DHMS. If we
assess the evolution of surface seawater and the mantle water mass, as suggested by the mantle wa-
ter content, we find that a slightly larger mantle water mass would occur with DHMS. However,
as suggested by the mantle water budget inferred from subduction modeling (e.g., Iwamori 2007),
the mantle water mass can be approximately 9 to 11 times larger than the present-day amount of
surface seawater based on the maximum water solubility along the slab geotherm computed using
a reference state of 60 Ma oceanic lithosphere at 1,573 K, which seems to be a realistic estimate
of water in the deep mantle. In this interpretation, the main water reservoir would be the mantle
transition zone, rather than the uppermost lower mantle corresponding to the DHMS solubility
because the slab geotherm would be sufficiently high that the DHMS may not host the water.
Currently, numerical modeling requires a water mass that is more than 10 times greater than the
present-day ocean mass, which is consistent with the possible mantle water content inferred by
Iwamori (2007), even at temperatures that are 100–200 K colder than the temperature expected

www.annualreviews.org • Dynamics in the Uppermost Lower Mantle 57


EA47CH03_Nakagawa ARjats.cls April 20, 2019 10:38

a Mantle water content

Water (wt%)
1

0.001
Annu. Rev. Earth Planet. Sci. 2019.47:41-66. Downloaded from www.annualreviews.org
Access provided by 186.183.193.242 on 05/03/20. For personal use only.

Surface seawater Mantle water mass


15
14
b Without DHMS c Without DHMS
Water content (OMs)

12 With DHMS With DHMS


10
8
6
4
2
0
0 0.5 1.0 1.5 2.0 2.5 3.0 3.5 4.0 4.5 0 0.5 1.0 1.5 2.0 2.5 3.0 3.5 4.0 4.5
Time (Ga) Time (Ga)
Figure 10
The lifetime of surface seawater in a plate–mantle dynamics system in numerical mantle convection simulations. (a) Mantle water
content field (left) without and (right) with dense hydrous magnesium silicate (DHMS). (b) Mass of surface seawater as a function of
time. (c) Mantle water mass profile. The vertical axes of panels b and c show the water mass normalized by the seawater mass at the
present day (1.4 × 1021 kg). Figure adapted from Nakagawa et al. (2018).

from Iwamori’s slab geotherm computation. Thus, water can be transported into the uppermost
lower mantle so that the DHMS can absorb this water. However, in current numerical modeling,
only the rheological properties of olivine are applied because few data for other hydrous minerals
have been provided by the mineral physics community. It is expected that obtaining more data
about the rheological properties of hydrous mantle minerals, particularly those of the uppermost
lower mantle minerals, will improve our current understanding of the amount of water in the
mantle. Nakagawa et al. (2018) also seem to include the huge amount of water on the surface in
the early Earth stage, but the initial amount of water in the mantle may not be very sensitive to
the mantle water mass at the present day because the initial amount of water in the deep mantle
should be regulated by the water solubility map (Nakagawa & Spiegelman 2017).

5. SUMMARY AND SYNTHESIS


In this review, we describe the current status of the numerical modeling of mantle dynamics in
the mantle transition zone and uppermost lower mantle, with a particular focus on the dynamics

58 Nakagawa • Nakakuki
EA47CH03_Nakagawa ARjats.cls April 20, 2019 10:38

of slab stagnation and penetration, as well as the mantle water cycle. Numerical modeling is able
to incorporate insights from physical and chemical interpretations of experimental data provided
by the mineral physics community, as well as from seismic imaging. However, further improve-
ments in numerical modeling techniques are still required to reproduce the schematic interpre-
tations provided by mineral physics measurements and observational data analyses. Additionally,
slab stagnation across the mantle transition zone and slab collapse due to mantle avalanche events
can provide significant information about the physics behind slab dynamics imaged by global to-
mography. The thermal and mechanical properties of subducting slabs are significant for revealing
these events in numerical modeling approaches.
The collapse of a stagnant slab is a key process in providing water to the deep lower mantle. The
uppermost lower mantle may be an additional reservoir of water transported via slab penetration.
The water solubility of the ambient lower mantle must be O(100) ppm to generate the hydrated
mantle transition zone predicted by the numerical modeling of hydrous mantle convection and
Annu. Rev. Earth Planet. Sci. 2019.47:41-66. Downloaded from www.annualreviews.org

slab stagnation. Such a transportation mechanism may be required to interpret the transportation
of hydrogen across the core–mantle boundary region because the effects of the solid solution of
Access provided by 186.183.193.242 on 05/03/20. For personal use only.

lower mantle minerals and iron hydrates may generate hydrogen in the lower mantle (Hu et al.
2016, Nishi et al. 2017).
Both rheological properties and the maximum water capacities of hydrous mantle minerals
play significant roles in the global-scale mantle water cycle, which may have a great impact on the
survival time of surface seawater. Again, further improvements of the numerical models of water
migration caused by dehydration reactions and the physical processes of hydration in the oceanic
crust are required. As shown in Figure 10b,c, the total water in the entire planetary system may
have more than 10 ocean masses. However, it is not clear that such a huge amount of water can be
transported in the planetary formation. Cosmochemical analysis, as well as N-body simulations of
planetary formation (Tian & Ida 2015), suggests that the proto-Earth may not have had so much
water in its interior (e.g., Albarede 2009), but some other analyses suggest that the early planet
could have had some amount of water in the deep interior (Marty 2012).
In addition, recent geochemical box models tracking the D/H ratio also suggest that the vol-
ume of water in an Earth-like planet may not be very large (Kurokawa et al. 2018). During the
atmospheric evolution associated with magma ocean solidification, an Earth-like planet may have
some amount of water (Hamano et al. 2013), consistent with the results of Marty (2012). Hamano
et al. (2013) and Marty (2012) also indicated a huge difference in the amount of water in the
early planetary interior among cosmochemical and geochemical modeling, atmospheric evolu-
tion modeling, and numerical mantle convection simulations provided in this review. To reduce
this gap, more careful and detailed investigations of the mechanism of water transport in planetary
formation and early atmospheric evolution should be done in the future using multidisciplinary
approaches (e.g., Precambrian geology, planetary formation theory, geo- and cosmochemical anal-
ysis). However, again, the initial amount of water in the deep mantle may not be very sensitive to
the mantle water mass at the present day, as suggested by numerical mantle convection simulations
(Nakagawa & Spiegelman 2017); the amount of water in the early Earth’s system is still uncertain
and should be determined by future investigations. The outgassing of volatiles via ridge volcan-
ism may change the atmospheric composition and affect the climate evolution and habitability of
rocky planets (e.g., Kadoya & Tajika 2015). However, a coupled model of climate evolution and
the deep planetary interior is still needed to incorporate the effects of plate motion, which could
help answer the fundamental questions of the formation of Earth-like habitable planets.
In the current view of the mantle water cycle, the mantle transition zone may be hydrated by
stagnation of subducted slabs with hydrous oceanic crust because the mantle transition zone may
have large solubility limits due to the properties of slab geotherm such as wadsleyite (Smyth &

www.annualreviews.org • Dynamics in the Uppermost Lower Mantle 59


EA47CH03_Nakagawa ARjats.cls April 20, 2019 10:38

Kawamoto 1997). When slab penetration occurs at some point, hydrated material can be trans-
formed into DHMS, which can still hold water. However, in the much deeper mantle, it is prob-
lematic to find a convincing mechanism of water storage after DHMS dissolves at approximately
80 GPa. Iron hydroxides may be stable at lower mantle conditions, but it is not clear if such miner-
als can host the water transported by subducting slabs into the deep mantle. Further investigations
of both mineral physics measurements and seismic data analyses are expected to include these in-
sights in numerical mantle convection simulations to reveal the hydrous mantle dynamics in the
uppermost lower mantle and mantle transition zone, as well as the possible mechanisms of water
transportation into the lowermost mantle. To resolve these issues, it is essential to determine the
amount of water the oceanic lithosphere can hold, which can be evaluated by analyzing the water–
rock interaction (serpentinization) associated with outer-rise faults near trenches (Hatakeyama
et al. 2017, Peacock 2001). Nakao et al. (2018) attempted to assess the model sensitivity of the
water content in the hydrous oceanic crust in a plate–mantle dynamics system, but there still re-
Annu. Rev. Earth Planet. Sci. 2019.47:41-66. Downloaded from www.annualreviews.org

mains room for evaluating the influence of the water flux transported into the deep mantle on
global-scale modeling. Moreover, for further understanding of the water circulation on the scale
Access provided by 186.183.193.242 on 05/03/20. For personal use only.

of the whole Earth (mantle and core), the partitioning of hydrogen between silicate melt and
molten iron should be addressed, as has been done by Okuchi (1997), although only at low pres-
sure. More detailed investigations of hydrogen in the deep Earth can be done with high-pressure
and high-temperature measurements of water- or hydrogen-bearing material in the near future.
It would be very interesting to determine whether the electrical conductivity in the mantle
transition zone and uppermost lower mantle is sensitive to the mantle water content (Khan &
Shankland 2012, Kuvshinov 2012); this is a topic that is not discussed in depth in this review.
It would be promising for the numerical modeling of mantle dynamics to obtain electrical con-
ductivity data from mantle minerals that are stable at the mantle transition zone and uppermost
lower mantle, as long as some potential issues with mineral physics could be resolved (Karato
2011, Yoshino & Katsura 2013). Deschamps & Khan (2016) attempted to compute the electrical
conductivity structure on the global scale with a probabilistic approach but still required various
improvements with a numerical modeling approach.
Finally, there are still many interesting topics on the first-order effects of global-scale mantle
dynamics incorporating water migration that require further study, as well as revealing the evolu-
tion of the whole planetary system (exosphere–interior interaction) with multidisciplinary efforts.

DISCLOSURE STATEMENT
The authors are not aware of any affiliations, memberships, funding, or financial holdings that
might be perceived as affecting the objectivity of this review.

ACKNOWLEDGMENTS
The authors thank Hikaru Iwamori, Atsuhi Nakao, Ryunosuke Yanagi, and Marc Spiegelman for
discussing the mantle water cycle; Hisayoshi Shimizu for providing information about the man-
tle electrical conductivity structure; Michio Tagawa, Masanori Kameyama, Yasuyuki Iwase, and
Shoichi Yoshioka for discussing the dynamics of stagnant slabs; and Shintaro Kadoya and Eiichi
Tajika for providing information about the interaction between deep planetary interior and surface
climate. The authors also thank Paul Tackley, Bruce Buffett, and Satoru Honda for their valuable
support. The authors are also appreciative of the reviewer, who greatly improved the original
manuscript.

60 Nakagawa • Nakakuki
EA47CH03_Nakagawa ARjats.cls April 20, 2019 10:38

LITERATURE CITED
Agrusta R, Goes S, van Hunen J. 2017. Subducting-slab transition-zone interaction: stagnation, penetration
and mode switches. Earth Planet. Sci. Lett. 464:10–23
Agrusta R, van Hunen J, Goes S. 2014. The effect of metastable pyroxene on the slab dynamics. Geophys. Res.
Lett. 41:8800–8
Albarede F. 2009. Volatile accretion history of the terrestrial planets and dynamic implications. Nature
461:1227–33
Angel RJ, Frost DJ, Ross NL, Hemley R. 2001. Stabilities and equations of state of dense hydrous magnesium
silicates. Phys. Earth Planet. Inter. 127:181–96
Arcey D, Tric E, Doin MP. 2005. Numerical simulations of subduction zones: effect of slab dehydration on
the mantle wedge dynamics. Phys. Earth Planet. Inter. 149:133–53
Arredondo KM, Billen MI. 2016. The effects of phase transitions and compositional layering in two-
dimensional kinematic models of subduction. J. Geodyn. 100:159–74
Annu. Rev. Earth Planet. Sci. 2019.47:41-66. Downloaded from www.annualreviews.org

Ballmer MD, Houser C, Hernlund JW, Wentzcovich R, Hirose K. 2017. Persistence of strong silica-enriched
domains in the Earth’s lower mantle. Nat. Geosci. 10:236–40
Ballmer MD, Schmerr NC, Nakagawa T, Ritsema J. 2015. Compositional mantle layering revealed by slab
Access provided by 186.183.193.242 on 05/03/20. For personal use only.

stagnation at ∼1000 km depth. Sci. Adv. 1:e1500815


Bercovici D, Karato S. 2003. Whole mantle convection and transition-zone water filter. Nature 425:39–44
Billen MI. 2010. Slab dynamics in the transition zone. Phys. Earth Planet. Inter. 183:296–308
Bolfan-Casanova N. 2005. Water in the Earth’s mantle. Mineral. Mag. 69:229–57
Chen C, Zhao D, Tian Y, Wu S, Hasegawa A, et al. 2017. Mantle transition zone, stagnant slab and intraplate
volcanism in Northeast Asia. Geophys. J. Int. 209:68–85
Christensen UR, Yuen DA. 1984. The interaction of a subducting slab with a chemical or phase boundary.
J. Geophys. Res. 89:4389–402
Cizkova H, Bina CR. 2013. Effects of mantle and subduction-interface rheologies on slab stagnation and
trench rollback. Earth Planet. Sci. Lett. 379:95–103
Cizkova H, van Hunen J, van den Berg AP, Vlaar NJ. 2002. The influence of rheological weakening and
yield stress on the interaction of slabs with the 670 km discontinuity. Earth Planet. Sci. Lett. 199:447–
57
Crowley JW, Gerault M, O’Connell RJ. 2011. On relative influence of heat and water transport on planetary
dynamics. Earth Planet. Sci. Lett. 310:380–88
Dannberg J, Heister T. 2016. Compressible magma/mantle dynamics: 3-D, adaptive simulations in ASPECT.
Geophys. J. Int. 207:1343–66
Deng J, Lee KKM. 2017. Viscosity jump in the lower mantle inferred from melting curves of ferropericlase.
Nat. Commun. 8:1997
Deschamps F, Khan A. 2016. Electrical conductivity as a constraint on lower mantle thermo-chemical struc-
ture. Earth Planet. Sci. Lett. 450:108–19
Deuss A. 2009. Global observations of mantle discontinuities using SS and PP precursors. Surv. Geophys.
30:301–26
Faccenda M. 2014. Water in the slab: a trilogy. Tectonophysics 614:1–30
Faccenda M, Dal Zilio L. 2017. The role of solid-solid phase transitions in mantle convection. Lithos 268–
71:198–224
Fei H, Wiedenback M, Yamazaki D, Katsura T. 2013. Small effect of water on upper-mantle rheology based
on silicon self-diffusion coefficients. Nature 298:213–15
Fei Y, van Orman J, Li J, van Westrenen W, Sanloup C, et al. 2004. Experimentally determined postspinel
transformation boundary in Mg2 SiO4 using MgO as an internal pressure standard and its geophysical
implications. J. Geophys. Res. 109:B02305
Fujita K, Ogawa M. 2013. A preliminary numerical study on water circulation in convecting mantle with
magmatism and tectonic plates. Phys. Earth Planet. Inter. 216:1–11
Fukao Y, Obayashi M. 2013. Subducted slabs stagnant above, penetrating through and trapped below the 660
km discontinuity. J. Geophys. Res. 118:5920–38

www.annualreviews.org • Dynamics in the Uppermost Lower Mantle 61


EA47CH03_Nakagawa ARjats.cls April 20, 2019 10:38

Fukao Y, Obayashi M, Nakakuki T, Deep Slab Proj. Group. 2009. Stagnant slab: a review. Annu. Rev. Earth
Planet. Sci. 37:19–46
Girard J, Amulele G, Farta R, Mohiuddin A, Karato S-I. 2016. Shear deformation of bridgmanite and magne-
siowüstite aggregates at lower mantle conditions. Science 351:144–47
Goes S, Agrusta R, van Hunen J, Garel F. 2017. Subduction-transition zone interaction: a review. Geosphere
13:644–64
Goes S, Capitanio FA, Morra G. 2008. Evidence for lower mantle slab penetration phases in plate motions.
Nature 451:981–84
Hamano K, Abe Y, Genda H. 2013. Emergence of two types of terrestrial planet on solidification of magma
ocean. Nature 497:607–10
Hatakeyama K, Katayama I, Hirauchi K, Michibayashi K. 2017. Mantle hydration along outer-rise faults in-
ferred from serpentinite permeability. Sci. Rep. 7:13870
He L. 2017. Wet plume atop of the flattening slab: insight into intraplate volcanism in East Asia. Phys. Earth
Planet. Inter. 269:29–39
Annu. Rev. Earth Planet. Sci. 2019.47:41-66. Downloaded from www.annualreviews.org

Hirose K. 2002. Phase transitions in pyrolitic mantle around 670-km depth: implications for upwelling of
plumes from the lower mantle. J. Geophys. Res. 107:B42078
Access provided by 186.183.193.242 on 05/03/20. For personal use only.

Hirth G, Kohlstedt DL. 1996. Water in the oceanic mantle: implications for rheology, melt extraction, and
the evolution of the lithosphere. Earth Planet. Sci. Lett. 144:93–108
Honda S. 2016. Slab stagnation and detachment under northeast China. Tectonophysics 671:127–38
Honda S. 2017. Geodynamic modeling of the subduction zone around the Japanese islands. Monogr. Environ.
Earth Planets 5:35–62
Honda S, Yuen DA, Balachandar S, Reuteler D. 1993. Three-dimensional instabilities of mantle convection
with multiple phase transitions. Science 259:1308–11
Horiuchi S, Iwamori H. 2016. A consistent model for fluid distribution, viscosity distribution, and flow-thermal
structure in subduction zone. J. Geophys. Res. 121:3238–60
Houser C. 2016. Global seismic data reveal little water in the mantle transition zone. Earth Planet. Sci. Lett.
448:94–101
Houser C, Williams Q. 2010. Reconciling Pacific 410 and 660 km discontinuity topography, transition zone
shear velocity patterns, and mantle phase transitions. Earth Planet. Sci. Lett. 296:187–97
Hu Q, Kim DY, Yang W, Yang L, Meng Y, et al. 2016. FeO2 and FeOOH under deep lower-mantle conditions
and Earth’s oxygen-hydrogen cycles. Nature 534:241–44
Ichiki M, Baba K, Obayashi M, Utada H. 2006. Water content and geotherm in the upper mantle above the
stagnant slab: interpretation of electrical conductivity and seismic P-wave velocity models. Phys. Earth
Planet. Inter. 155:1–15
Irifune T, Ringwood AE. 1993. Phase transformations in subducted oceanic crust and buoyancy relationships
at depths of 600–800 km in the mantle. Earth Planet. Sci. Lett. 117:101–10
Ito E, Takahashi E. 1989. Postspinel transformation in the system Mg2 SiO4 -Fe2 SiO4 and some geophysical
implications. J. Geophys. Res. 94:10637–46
Iwamori H. 2004. Phase relations of peridotites under H2 O-saturated conditions and ability of subducting
plates for transportation of H2 O. Earth Planet. Sci. Lett. 227:57–71
Iwamori H. 2007. Transportation of H2 O beneath the Japan arcs and its implications for global water circu-
lation. Chem. Geol. 239:182–98
Iwamori H, Nakakuki T. 2013. Fluid processes in subduction zones and water transport to the deep mantle.
In Physics and Chemistry of the Deep Earth, ed. S-I Karato, pp. 372–91. Hoboken, NJ: Wiley
Jacobsen SD, Jiang F, Mao Z, Duffy TS, Smyth JR, et al. 2008. Effects of hydration on the elastic properties
of olivine. Geophys. Res. Lett. 35:L14303
Kadoya S, Tajika E. 2015. Evolutionary climate tracks of Earth-like planets. Astrophys. J. Lett. 815:L7
Kameyama M, Nishioka R. 2012. Generation of ascending flow in the Big Mantle Wedge (BMW) beneath
Northeast Asia induced by retreat and stagnation of subducted slab. Geophys. Res. Lett. 39:L10309
Kaneshima S. 2009. Seismic scatters at the shallowest lower mantle beneath subducted slabs. Earth Planet. Sci.
Lett. 286:304–15

62 Nakagawa • Nakakuki
EA47CH03_Nakagawa ARjats.cls April 20, 2019 10:38

Karato S. 2011. Water distribution across the mantle transition zone and its implications for global material
circulation. Earth Planet. Sci. Lett. 301:413–23
Karato S, Bercovici D, Leahy G, Richard G, Jing Z. 2006. The transition-zone water filter model for global
material circulation: Where do we stand? In Earth’s Deep Water Cycle, ed. SD Jacobsen, S Van Der Lee,
pp. 289–313. Washington, DC: Am. Geophys. Union
Karato S, Riedel MR, Yuen DA. 2001. Rheological structure and deformation of subducted slabs in the man-
tle transition zone: implications for mantle circulation and deep earthquakes. Phys. Earth Planet. Inter.
127:83–108
Karato S-I, Wu P. 1993. Rheology of the upper mantle: a synthesis. Science 260:771–78
Katsura T, Yamada H, Shinmei T, Kudo A, Ono S, et al. 2003. Post-spinel transition in Mg2 SiO4 determined
by high P-T in situ X-ray diffractometry. Phys. Earth Planet. Inter. 136:11–24
Katz RF, Spiegelman M, Langmuir CH. 2003. A new parameterization of hydrous mantle melting. Geochem.
Geophys. Geosyst. 4:1073
Kaufmann G, Lambeck K. 2000. Mantle dynamics, postglacial rebound and radial viscosity profile. Phys. Earth
Annu. Rev. Earth Planet. Sci. 2019.47:41-66. Downloaded from www.annualreviews.org

Planet. Inter. 121:301–24


Kawakatsu H, Watada S. 2007. Seismic evidence for deep-water transportation in the mantle. Science 316:1468–
Access provided by 186.183.193.242 on 05/03/20. For personal use only.

71
Kawamoto T, Holloway JR. 1997. Melting temperature and partial melt chemistry of H2 O-saturated peri-
dotite to 11 gigapascals. Science 276:240–43
Khan A, Shankland TJ. 2012. A geophysical perspective on mantle water content and melting: inverting elec-
tromagnetic sounding data using laboratory-based electrical conductivity profiles. Earth Planet. Sci. Lett.
317:27–43
King SD. 1995. Models of mantle viscosity. In Mineral Physics and Crystallography: A Handbook of Physical Con-
stants, Vol. 2, ed. TJ Ahrens, pp. 227–36. Washington, DC: Am. Geophys. Union
Komabayashi T, Omori S, Maruyama S. 2004. Petrogenetic grid in the system MgO-SiO2 -H2 O up to
30 GPa, 1600C: applications to hydrous peridotite subducting into the Earth’s deep interior. J. Geophys.
Res. 109:B03206
Komacek TS, Abbot DS. 2016. Effect of surface-mantle water exchange parameterizations on exoplanet ocean
depths. Astrophys. J. 832:54
Korenaga J. 2011. Thermal evolution with a hydrating mantle and the initiation of plate tectonics in the early
Earth. J. Geophys. Res. 116:B12403
Korenaga J, Karato S-I. 2008. A new analysis of experimental data on olivine rheology. J. Geophys. Res.
113:B02403
Kubo T, Kaneshima S, Torii Y, Yoshioka S. 2009. Seismological and experimental constraints on metastable
phase transformations and rheology of the Mariana slab. Earth Planet. Sci. Lett. 287:12–23
Kurokawa H, Foriel J, Laneuville M, Houser C, Usui T. 2018. Subduction and atmospheric escape of Earth’s
seawater constrained by hydrogen isotopes. Earth Planet. Sci. Lett. 497:149–60
Kuvshinov AV. 2012. Deep electromagnetic studies from land, sea and space: progress status in the past
10 years. Surv. Geophys. 33:169–209
Li Z-XA, Lee C-TA, Peslier A, Lenerdic A, Mackwell SJ. 2008. Water contents in mantle xenoliths from the
Colorado Plateau and vicinity: implications for the rheology and hydrogen-induced thinning continental
lithosphere. J. Geophys. Res. 113:B09210
Marty B. 2012. The origins and concentrations of water, carbon, nitrogen and noble gases on Earth. Earth
Planet. Sci. Lett. 313–14:56–66
Mei S, Kohlstedt DL. 2000. Influence of water on plastic deformation of olivine aggregates: 1. Diffusion creep
regime. J. Geophys. Res. 105:21457–69
Mitrovica JX, Forte AM. 1997. Radial profile of mantle viscosity: results from the joint inversion of convection
and postglacial rebound observables. J. Geophys. Res. 102:2751–69
Murakami M, Hirose K, Kawamura K, Sata N, Ohishi Y. 2004. Post-perovskite phase transition in MgSiO3 .
Science 304:855–58
Murakami M, Hirose K, Yurimoto H, Nakashima S, Takafuji N. 2002. Water in Earth’s lower mantle. Science
295:1885–87

www.annualreviews.org • Dynamics in the Uppermost Lower Mantle 63


EA47CH03_Nakagawa ARjats.cls April 20, 2019 10:38

Nakada M, Okuno J. 2016. Inference of mantle viscosity for depth resolutions of GIA observations. Geophys.
J. Int. 207:719–40
Nakagawa T. 2017. On the numerical modeling of the deep mantle water cycle in global-scale mantle dynamics:
the effects of the water solubility limit of lower mantle minerals. J. Earth Sci. 28:563–77
Nakagawa T, Buffett BA. 2005. Mass transport mechanism between the upper and lower mantle in numerical
simulations of thermochemical mantle convection with multi-component phase changes. Earth Planet.
Sci. Lett. 230:11–27
Nakagawa T, Iwamori H. 2017. Long-term stability of plate-like behavior in hydrous mantle convection and
water absorption into the deep mantle. J. Geophys. Res. Solid Earth 122:8431–45
Nakagawa T, Iwamori H, Yanagi R, Nakao A. 2018. On the evolution of the water ocean in the plate-mantle
system. Prog. Earth Planet. Sci. 5:51
Nakagawa T, Nakakuki T, Iwamori H. 2015. Water circulation and global mantle dynamics: insight from
numerical modeling. Geochem. Geophys. Geosyst. 16:1449–64
Nakagawa T, Spiegelman M. 2017. Global-scale water circulation in the Earth’s mantle: implications for the
Annu. Rev. Earth Planet. Sci. 2019.47:41-66. Downloaded from www.annualreviews.org

mantle water budget in the early Earth. Earth Planet. Sci. Lett. 464:189–99
Nakagawa T, Tackley PJ, Deschamps F, Connolly JAD. 2010. The influence of MORB and Harzburgite com-
Access provided by 186.183.193.242 on 05/03/20. For personal use only.

position on thermo-chemical mantle convection in a 3-D spherical shell with self-consistently calculated
mineral physics. Earth Planet. Sci. Lett. 296:403–12
Nakakuki T, Sato H, Fujimoto H. 1994. Interaction of the upwelling plume with the phase and chemical
boundaries at the 670 km discontinuity: effects of temperature-dependent viscosity. Earth Planet. Sci.
Lett. 121:369–84
Nakakuki T, Tagawa M, Iwase Y. 2010. Dynamical mechanism controlling formation and avalanche of a stag-
nant slab. Phys. Earth Planet. Inter. 183:309–20
Nakao A, Iwamori H, Nakakuki T. 2016. Effects of water transportation on subduction dynamics: roles of
viscosity and density reduction. Earth Planet. Sci. Lett. 454:178–91
Nakao A, Iwamori H, Nakakuki T, Suzuki YJ, Nakamura H. 2018. Roles of hydrous lithospheric mantle in
deep water transportation and subduction dynamics. Geophys. Res. Lett. 45:5336–43
Nishi M, Irifune T, Tsuchiya J, Nishihara Y, Fujino K, Higo Y. 2014. Stability of hydrous silicate at high
pressures and water transport to the deep lower mantle. Nat. Geosci. 7:224–27
Nishi M, Kuwayama Y, Tsuchiya J, Tsuchiya T. 2017. The pyrite-type high-pressure form of FeOOH. Nature
547:205–8
Oganov AR, Ono S. 2004. Theoretical and experimental evidence for a post-perovskite phase of MgSiO3 in
Earth’s D layer. Nature 430:445–48
Ohira I, Ohtani E, Sakai T, Miyahara M, Hirao N, et al. 2014. Stabiliy of a hydrous δ-phase, AlOOG-
MgSiO2 (OH)2 , and a mechanism for water transport into the base of lower mantle. Earth Planet. Sci.
Lett. 401:12–17
Okuchi T. 1997. Hydrogen partitioning into molten iron at high pressure: implications for Earth’s core. Science
278:1781–84
Ono S, Ito E, Katsura T. 2001. Mineralogy of subducted basaltic crust (MORB) from 25 to 37 GPa, and
chemical heterogeneity of the lower mantle. Earth Planet. Sci. Lett. 190:57–63
Panero WR, Pigott JS, Reaman DM, Kabbes JE, Liu Z. 2015. Dry (Mg,Fe)SiO3 perovskite in the Earth’s
lower mantle. J. Geophys. Res. Solid Earth 120:894–908
Peacock SM. 2001. Are the lower planes of double seismic zones caused by serpentine dehydration in sub-
ducting oceanic mantle? Geology 29:299–302
Pearson DG, Brenker FE Nestola F, McNeill J, Nasdala L, et al. 2014. Hydrous mantle transition zone indi-
cated by ringwoodite included within diamond. Nature 507:221–24
Richard GC, Bercovici D. 2009. Water-induced convection in the Earth’s mantle transition zone. J. Geophys.
Res. 114:B01205
Richard GC, Iwamori H. 2010. Stagnant slab, wet plumes and Cenozoic volcanism in East Asia. Phys. Earth
Planet. Inter. 183:280–87
Richard GC, Monneraeu M, Ingrin J. 2002. Is the transition zone an empty water reservoir? Influence from
numerical model of mantle dynamics. Earth Planet. Sci. Lett. 205:37–51

64 Nakagawa • Nakakuki
EA47CH03_Nakagawa ARjats.cls April 20, 2019 10:38

Rohrbach A, Schmidt MW. 2011. Redoc freezing and melting in the Earth’s deep mantle resulting from
carbon-iron redox coupling. Nature 472:209–12
Rubie DC. 1984. The olivine-spinel transformation and the rheology of subducting lithosphere. Nature
308:505–8
Rudolph ML, Lekic V, Lithgow-Bertelloni C. 2015. Viscosity jump in Earth’s mid-mantle. Science 350:1349–
52
Rüpke LH, Morgan JP, Hort M, Connolly JAD. 2004. Serpentine and the subduction zone water cycle. Earth
Planet. Sci. Lett. 223:17–34
Sandu C, Lenardic A, McGovern P. 2011. The effects of deep water cycling on planetary thermal evolution.
J. Geophys. Res. 116:B12404
Schmandt B, Jacobsen SD, Becker TW, Liu Z, Dueker KG. 2014. Dehydration melting at the top of the lower
mantle. Science 344:1265–68
Schmeling H, Monz R, Rubie DC. 1999. The influence of olivine metastability on the dynamics of subduction.
Earth Planet. Sci. Lett. 165:55–66
Annu. Rev. Earth Planet. Sci. 2019.47:41-66. Downloaded from www.annualreviews.org

Scire A, Zandt G, Beck S, Long M, Wagner L. 2017. The deforming Nazca slab in the mantle transition zone
and lower mantle: constraints from teleseismic tomography on the deeply subducted slab between 6S
Access provided by 186.183.193.242 on 05/03/20. For personal use only.

and 32S. Geosphere 13:665–80


Smyth JR, Kawamoto T. 1997. Wadsleyite II: a new high pressure hydrous phase in the peridotite-H2 O system.
Earth Planet. Sci. Lett. 146:E9–16
Solheim LP, Peltier WR. 1994. Avalanche effects in phase transition modulated thermal convection, a model
of Earth’s mantle. J. Geophys. Res. 99:6997–7018
Stixrude L, Lithgow-Bertelloni C. 2012. Geophysics of chemical heterogeneity in the mantle. Annu. Rev. Earth
Planet. Sci. 40:569–95
Suetsugu D, Inoue T, Yamada A, Zhao D, Obayashi M. 2006. Towards mapping the three-dimensional distri-
bution of water in the transition zone from P-velocity tomography and 660-km discontinuity depths. In
Earth’s Deep Water Cycle, ed. SD Jacobsen, S Van Der Lee, pp. 237–49. Washington, DC: Am. Geophys.
Union
Tackley PJ. 1995. On the penetration of an endothermic phase transition by upwellings and downwellings.
J. Geophys. Res. 100:15477–88
Tackley PJ. 1996. On the ability of phase transitions and viscosity layering to induce long wavelength hetero-
geneity in the mantle. Geophys. Res. Lett. 23:1985–88
Tackley PJ. 2000. Mantle convection and plate tectonics: towards an integrated physical and chemical theory.
Science 288:2002–7
Tackley PJ, Stevenson DJ, Glatzmaier GA, Schubert G. 1993. Effects of an endothermic phase transition at
670 km depth in a spherical model of convection in the Earth’s mantle. Nature 361:699–704
Tackley PJ, Xie S, Nakagawa T, Hernlund JW. 2005. Numerical and laboratory studies of mantle convec-
tion: philosophy, accomplishments and thermo-chemical structure and evolution. In Earth’s Deep Man-
tle: Structure, Composition, and Evolution, ed. RD Van Der Hilst, JD Bass, J Matas, J Tramper, pp. 83–99.
Washington, DC: Am. Geophys. Union
Tagawa M, Nakakuki T, Tajima F. 2007. Dynamical modeling of trench retreat driven by the slab interaction
with the mantle transition zone. Earth Planets Space 59:65–74
Thoraval C, Richards MA. 1997. The geoid constraint in global geodynamics: viscosity structure, mantle
heterogeneity models and boundary conditions. Geophys. J. Int. 131:1–8
Tian F, Ida S. 2015. Water contents of Earth-mass planets around M dwarfs. Nat. Geosci. 8:177–80
Torii Y, Yoshioka S. 2007. Physical conditions producing slab stagnation: constraints of the Clapeyron slope,
mantle viscosity, trench retreat, and dip angles. Tectonophysics 445:200–9
van Keken PE, Hacker BR, Syracuse EM, Abers GA. 2011. Subduction factory: 4. Depth-dependent flux of
H2 O from subducting slabs worldwide. J. Geophys. Res. 116:B01401
Wang J, Sinogeikin SV, Inoue T, Bass JD. 2006. Elastic properties of hydrous ringwoodite at high-pressure
conditions. Geophys. Res. Lett. 33:L14308
Wang Z, Kusky TM, Capitanio FA. 2018. Water transportation ability of flat-lying slabs in the mantle transi-
tion zone and implications for craton destruction. Tectonophysics 723:95–106

www.annualreviews.org • Dynamics in the Uppermost Lower Mantle 65


EA47CH03_Nakagawa ARjats.cls April 20, 2019 10:38

Waszek L, Schmerr NC, Ballmer MD. 2018. Global observations of reflectors in the mid-mantle with impli-
cations for mantle structure and dynamics. Nat. Commun. 9:385
Wilson CR, Spiegelman M, van Keken PE, Hacker BR. 2014. Fluid flow in subduction zones: the role of solid
rheology and compaction pressure. Earth Planet. Sci. Lett. 401:261–74
Ye L, Li J, Tseng T, Yao Z. 2011. A stagnant slab in a water-bearing mantle transition zone beneath northeast
China: implications from regional SH waveform modelling. Geophys. J. Int. 186:706–10
Ye Y, Brown DA, Smyth JR, Panero WR, Jacobsen SD, et al. 2012. Compressibility and thermal expansivity
of hydrous ringwoodite with 2.5(3) wt% H2 O. Am. Mineral. 97:573–82
Yoshino T, Katsura T. 2013. Electrical conductivity of mantle minerals: role of water in conductive anomalies.
Annu. Rev. Earth Planet. Sci. 41:605–28
Yoshioka S, Naganoda A. 2010. Effects of trench migration on fall of stagnant slabs into the lower mantle.
Phys. Earth Planet. Inter. 183:321–29
Annu. Rev. Earth Planet. Sci. 2019.47:41-66. Downloaded from www.annualreviews.org
Access provided by 186.183.193.242 on 05/03/20. For personal use only.

66 Nakagawa • Nakakuki
EA47_FrontMatter ARI 26 April 2019 11:24

Annual Review
of Earth and
Planetary Sciences

Volume 47, 2019 Contents

Big Time
Paul F. Hoffman p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 1
Unanticipated Uses of the Global Positioning System
Annu. Rev. Earth Planet. Sci. 2019.47:41-66. Downloaded from www.annualreviews.org

Kristine M. Larson p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p19


Access provided by 186.183.193.242 on 05/03/20. For personal use only.

Dynamics in the Uppermost Lower Mantle: Insights into the Deep


Mantle Water Cycle Based on the Numerical Modeling of
Subducted Slabs and Global-Scale Mantle Dynamics
Takashi Nakagawa and Tomoeki Nakakuki p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p41
Atmospheric Escape and the Evolution of Close-In Exoplanets
James E. Owen p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p67
The Sedimentary Cycle on Early Mars
Scott M. McLennan, John P. Grotzinger, Joel A. Hurowitz, and Nicholas J. Tosca p p p p p p91
New Horizons Observations of the Atmosphere of Pluto
G. Randall Gladstone and Leslie A. Young p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 119
The Compositional Diversity of Low-Mass Exoplanets
Daniel Jontof-Hutter p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 141
Destruction of the North China Craton in the Mesozoic
Fu-Yuan Wu, Jin-Hui Yang, Yi-Gang Xu, Simon A. Wilde,
and Richard J. Walker p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 173
Seawater Chemistry Through Phanerozoic Time
Alexandra V. Turchyn and Donald J. DePaolo p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 197
Global Patterns of Carbon Dioxide Variability from Satellite
Observations
Xun Jiang and Yuk L. Yung p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 225
Permeability of Clays and Shales
C.E. Neuzil p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 247
Flood Basalts and Mass Extinctions
Matthew E. Clapham and Paul R. Renne p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 275
Repeating Earthquakes
Naoki Uchida and Roland Bürgmann p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 305

viii
EA47_FrontMatter ARI 26 April 2019 11:24

Soil Functions: Connecting Earth’s Critical Zone


Steven A. Banwart, Nikolaos P. Nikolaidis, Yong-Guan Zhu, Caroline L. Peacock,
and Donald L. Sparks p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 333
Earthquake Early Warning: Advances, Scientific Challenges,
and Societal Needs
Richard M. Allen and Diego Melgar p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 361
Noble Gases: A Record of Earth’s Evolution and Mantle Dynamics
Sujoy Mukhopadhyay and Rita Parai p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 389
Supraglacial Streams and Rivers
Lincoln H Pitcher and Laurence C. Smith p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 421
Annu. Rev. Earth Planet. Sci. 2019.47:41-66. Downloaded from www.annualreviews.org

Isotopes in the Water Cycle: Regional- to Global-Scale Patterns and


Applications
Access provided by 186.183.193.242 on 05/03/20. For personal use only.

Gabriel J. Bowen, Zhongyin Cai, Richard P. Fiorella, and Annie L. Putman p p p p p p p p p p 453
Marsh Processes and Their Response to Climate Change
and Sea-Level Rise
Duncan M. FitzGerald and Zoe Hughes p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 481
The Mesozoic Biogeographic History of Gondwanan Terrestrial
Vertebrates: Insights from Madagascar’s Fossil Record
David W. Krause, Joseph J.W. Sertich, Patrick M. O’Connor,
Kristina Curry Rogers, and Raymond R. Rogers p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 519
Droughts, Wildfires, and Forest Carbon Cycling: A Pantropical
Synthesis
Paulo M. Brando, Lucas Paolucci, Caroline C. Ummenhofer, Elsa M. Ordway,
Henrik Hartmann, Megan E. Cattau, Ludmila Rattis, Vincent Medjibe,
Michael T. Coe, and Jennifer Balch p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 555
Exoplanet Clouds
Christiane Helling p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 583

Errata

An online log of corrections to Annual Review of Earth and Planetary Sciences articles
may be found at http://www.annualreviews.org/errata/earth

Contents ix

Potrebbero piacerti anche