83
Tracking Slabs Beneath Northeastern Pacific Subduction Zones Yu Jeffrey Gu University of Alberta, Department of Physics, CEB 348-D, Edmonton, AB, Canada, T6G 2G7. E-mail: [email protected] Phone: 1 780 492 2292 Fax: 1 780 492 0714 Ahmet Okeler University of Alberta, Department of Physics, CEB 456, Edmonton, AB, Canada, T6G 2G7. E-mail: [email protected] Phone: 1 780 492 4125 Fax: 1 780 492 0714 Ryan Schultz 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19

University of Alberta › ~ygu › publications › japan_pa… · Web viewTracking Slabs Beneath Northeastern Pacific Subduction Zones Yu Jeffrey Gu University of Alberta, Department

  • Upload
    others

  • View
    2

  • Download
    0

Embed Size (px)

Citation preview

Page 1: University of Alberta › ~ygu › publications › japan_pa… · Web viewTracking Slabs Beneath Northeastern Pacific Subduction Zones Yu Jeffrey Gu University of Alberta, Department

Tracking Slabs Beneath Northeastern Pacific Subduction Zones

Yu Jeffrey Gu

University of Alberta, Department of Physics, CEB 348-D, Edmonton, AB, Canada, T6G 2G7.

E-mail: [email protected]

Phone: 1 780 492 2292

Fax: 1 780 492 0714

Ahmet Okeler

University of Alberta, Department of Physics, CEB 456, Edmonton, AB, Canada, T6G 2G7.

E-mail: [email protected]

Phone: 1 780 492 4125

Fax: 1 780 492 0714

Ryan Schultz

University of Alberta, Department of Physics, CEB 456, Edmonton, AB, Canada, T6G 2G7.

E-mail: [email protected]

Phone: 1 780 492 4125

Fax: 1 780 492 0714

1

2

3

4

5

6

7

8

9

10

11

12

13

14

15

16

17

18

19

20

21

22

23

Page 2: University of Alberta › ~ygu › publications › japan_pa… · Web viewTracking Slabs Beneath Northeastern Pacific Subduction Zones Yu Jeffrey Gu University of Alberta, Department

Abstract

Illuminating major thermal and/or compositional variations in Earth's mantle based on reflected

seismic waves is analogous to “motion tracking” in animation cinematography. Signals analyzed

by both approaches are sensitive to strong gradients in material properties and, with proper

treatments, can be used to decipher the shape or movements of the enclosed mass. In the same

spirit, this study utilizes the amplitudes of bottom-side reflected shear waves to provide first-

order constraints on the geometry and kinematics of subducted oceanic crust and lithosphere

beneath the northwestern Pacific subduction zones. The high-resolution, depth-migrated

reflection amplitudes shows large, ~1000 km wide depressions on the 660-km seismic

discontinuity, extending from the Japan sea to eastern China. The 410-km seismic discontinuity

is locally elevated by ~15 km on the oceanside of the Japan trench, where a sharp change of

transition zone thickness infers a mantle temperature increase over XX deg C. The 410-km

seismic discontinuity is locally elevated by ~15 km east of the Wadati-Benioff zone, within

which reflection amplitude drops off significantly. We further identify a strong reflector at ~530

km depth with a reflection amplitude exceeding 5% of SS amplitude. The strength of this

anomaly increases depressed with ‘avalanching’ the lower mantle west of the Hokkaido corner.

Strong correlations between the reflectivity structure and seismic velocity suggest: (1) high-

amplitude reflections generally occurs near the edges of major seismic anomalies due to strong

shear wave focusing effect, (2) ‘gaps’ in the reflection amplitudes of the 410- and 660-km

seismic discontinuities are associated with substantial topography and major mass/heat fluxes.,

and (3). The presence this reflectors residual plume(s) in this region. UNFINISHED, will work

on last.

1

2

3

4

5

6

7

8

9

10

11

12

13

14

15

16

17

18

19

20

21

22

23

Page 3: University of Alberta › ~ygu › publications › japan_pa… · Web viewTracking Slabs Beneath Northeastern Pacific Subduction Zones Yu Jeffrey Gu University of Alberta, Department

1. Introduction

The convergent boundary between the Pacific, Amurian, and North American plates represents

one of the fastest destruction zones of old oceanic domains. The subduction process in this re-

gion initiated during the Cretaceous times (~65-140 Ma ago) (Northrup et al., 1995; Tonegawa et

al., 2006; Zhu et al., 2010) and continues to accommodate the differential motions between the

Pacific, Eurasia, and North American plates. The deposition of old oceanic lithosphere at the

present rate of 8-9.5 cm/yr (DeMets et al, 1990; Seno et al., 1996; Bird, 2003) not only directly

influences the surrounding mantle temperature and/or mineralogy.

The morphology and spatial extent of subducted oceanic lithosphere (for short, ‘slab’) beneath

the northwestern Pacific margin have long been investigated. Among the various data types and

approaches, seismic tomography of body waves has been the most effective in constraining

details of slab geometry and surrounding mantle conditions in this region (e.g., van der Hilst et

al., 1991, 1997; Fukao, 1992; Bijwaard et al., 1998; Fukao et al., 2001; Obayashi et al. 2006;

Huang and Zhao, 2006; Zhao and Ohtani, 2009; Li and van der Hilst, 2010). Well-defined zones

of above-average P and S wave speeds have been identified along the Wadati-Benioff zone and

within the upper mantle transition zone near Korea and eastern China (e.g., Jordan, 1977; van der

Hilst et al., 1997; Widiyantoro et al., 1997; Bijwaard et al., 1998; K´arason and van der Hilst,

2000; Fukao et al., 2001; Gorbatov et al., 2000; Gorbatov and Kennet, 2002; Lebedev and Nolet,

2003; Zhao, 2004; Obayashi et al., 2006; Huang and Zhao, 2006; Fukao et al., 2009; Zhao and

Ohtani, 2009; Li and van der Hilst, 2010). The non-geometrical shape of the high-velocity zones

have inspired discussions of slab deflection toward the horizontal, which is generally referred to

1

2

3

4

5

6

7

8

9

10

11

12

13

14

15

16

17

18

19

20

21

22

23

Page 4: University of Alberta › ~ygu › publications › japan_pa… · Web viewTracking Slabs Beneath Northeastern Pacific Subduction Zones Yu Jeffrey Gu University of Alberta, Department

as ‘stagnation’ (Fukao et al., 1992; Fukao et al., 2001), and possible extension into the lower

mantle (see Fukao et al., 2001, 2009 for detailed reviews). The length of the flattened part of the

slab can be as large as 800-1000 km (Huang and Zhao, 2006; Obayashi et al., 2006; Fukao et al.,

2009), at least half of which can be reproduced numerically with proper treatments of trench

migration and rollback rates (). Low-velocity structures such as arc volcanism and/or

decompressional melting of stagnant slabs (Lebedev and Nolet, 2003; Zhao, 2004; Priestley et

al., 2006; Obayashi et al., 2006; Zhao and Ohtani, 2009; An et al., 2009; Wang et al., 2009; Duan

et al., 2009; Zhao et al., 2009; Li and van der Hilst, 2010; Feng and An, 2010), or hot thermal

plume(s) (Miyashiro, 1986; Ichiki et al., 2006; Zou et al., 2008; Zhao and Ohtani, 2009; Duan et

al., 2009), further underscores the wide range of dynamical processes beneath this region. These

low- and high-velocity heterogeneities can cause strong gradients in mantle temperature and/or

composition surrounding the convergent plate boundary zones

In comparison with seismic tomography, which is highly effective in resolving ‘smooth’

variations, the amplitudes and arrival-times of body waves reflected and converted at mantle

depths are more sensitive to sharp changes in rock elastic properties (Zheng et al., 2007).

Correlations between velocity and reflectivity (Shearer and Masters, 1992; Flanagan and Shearer,

1998; Li et al., 2000; Shen et al., 2008) offer greater constraints on slab geometry and dynamics

than either approach alone. For this reason, the temperature-dependent depressions on the 660-

km seismic discontinuity by 15-60 km (Shearer and Masters, 1992; Benz and Vidale, 1992; Bina

and Helfrich? Helfrich and Bina?? Li et al., 2000; Niu et al., 2005; Tonegawa et al., 2005; Shen

et al., 2008; Tauzin et al., 200??; Lawrence and Shearer, 2006; Houser et al., 2008) have been

widely cited as evidence of stagnating and ponding slab beneath the northwestern Pacific

1

2

3

4

5

6

7

8

9

10

11

12

13

14

15

16

17

18

19

20

21

22

23

Page 5: University of Alberta › ~ygu › publications › japan_pa… · Web viewTracking Slabs Beneath Northeastern Pacific Subduction Zones Yu Jeffrey Gu University of Alberta, Department

collision zone. The resolutions of these seismic surveys are, however, hampered by the

restrictive source-receiver distributions of converted phases and the large averaging radii in

global analyses of secondary reflected waves also known as ‘SS precursors’. In particular, while

a pioneering study of the latter phase (Shearer and Masters, 1992) provided evidence of

stagnating slab beneath the northwestern Pacific region nearly 20 years ago, further usage of

these phases in constraining detailed slab geometry and kinematics was debated (Neele et al.,

1997; Shearer et al., 1999). Discussions of the correlations between mantle reflectivity inferred

from SS precursors and seismic velocities/mantle mineralogy near subduction zones mainly

focused on broad length scales and remained qualitative (e.g., Gu et al., 2003; Lawrence and

Shearer, 2006; Houser et al., 2008).

This study analyzes a large regional dataset of SS precursors using novel processing techniques

to improve the resolution on the seismic reflectivity structure beneath the northwestern Pacific

region (Fig. 1A). The dense regional data coverage enables pre-stack depth migration that

positions weak SS precursor amplitudes at the appropriate reflection depths and locations. By

correlating reflection amplitude variations with wave speeds, we aim to provide a self-consistent,

three-dimensional (3D) snapshot of mantle reflectivity structure and deformation near the

northwestern segment of the Pacific Ocean basin. For brevity we will hereon refer to the upper

mantle transition zone as MTZ and the 410-km, 520-km and 660-km discontinuities as the 410,

520 and 660, respectively.

2. Data and method

SS precursors are a proven means for determining the depths of mantle reflectors (e.g., Shearer

and Masters, 1992; Shearer, 1993; Gossler and Kind, 1996; Gu et al., 1998; Deuss and

1

2

3

4

5

6

7

8

9

10

11

12

13

14

15

16

17

18

19

20

21

22

23

24

Page 6: University of Alberta › ~ygu › publications › japan_pa… · Web viewTracking Slabs Beneath Northeastern Pacific Subduction Zones Yu Jeffrey Gu University of Alberta, Department

Woodhouse, 2002; Flanagan and Shearer, 1998; Gu and Dziewonski, 2002; Schmerr and

Garnero, 2007; Lawrence and Shearer, 2007; Houser et al., 200XX; Rychert and Shearer, ??).

Their strong sensitivities to the reflection depth and interfacial impedance contrast beneath mid

points (see Fig. 1A), coupled with their strong sensitivity to structures away from the source and

station locations, are ideal for mapping mantle reflectivity at both global and regional scales.

We utilize all available broadband, high-gain recordings of earthquakes that took place prior to

2008. This data set is currently managed by the IRIS Data Management Center and highlights

significant efforts from GDSN, IRIS, GEOSCOPE and several other temporary deployments.

Only data from shallow events (<75 km) with magnitude (Mw) grater than 5.0 are selected for

this undertaking. The former criterion minimizes the effect of depth phase, and the subjective

magnitude cutoff ensure that source mechanism solutions are available from the Global Centroid

Moment Tensor (GCMT) project (Dziewonski and Woodhouse, 1983) for accurate computations

of PREM (Dziewonski and Anderson, 1981) synthetic seismograms. We further restrict the

epicenter distance range to 100°-160° to minimize known waveform interferences from topside

reflection sdsS and ScS precursors ScSdScS, where d denotes a discontinuity (Schmerr and

Garnero, 2007). After applying a Butterworth band-pass filter with corner periods at 12 s and 75

s to the selected data traces, we impose a signal-to-noise ratio (SNR) criterion as the ratio

between the maximum absolute amplitude of the SS and noise. The selected signal and noise

windows are (-20 sec, 60 sec) and (-170 sec, -80 sec), respectively, relative to the predicted

arrival time of SS based on PREM (Dziewonski and Anderson, 1981). All records with SNR

lower than 3.0 are automatically rejected.

1

2

3

4

5

6

7

8

9

10

11

12

13

14

15

16

17

18

19

20

21

22

23

Page 7: University of Alberta › ~ygu › publications › japan_pa… · Web viewTracking Slabs Beneath Northeastern Pacific Subduction Zones Yu Jeffrey Gu University of Alberta, Department

The selected transverse-component seismograms are subsequently aligned on the first major

swing of SS phase with the aid of the corresponding synthetic seismograms. As the last step of

pre-processing, we apply time shifts by the theoretical SS and S520S times through PREM

(Dziewonski and Anderson, 1981) to account for crustal (Bassin et al., 2000) and topographical

(ETOPO5 data base) variations. Since our main focus is the upper mantle transition zone, the

approximation based on SS-S520S represents an effective compromise between the 410 and 660

and may introduce an error of 3-5 km for the depth estimation of reflectors hundreds of

kilometers away from the MTZ. Generally, these model assumptions have greater impacts on the

differential times, hence reflection depths, than on the amplitudes of SS precursors (e.g., Gu et

al., 2003).

A time-to-depth migration approach, which has been previously applied to P-to-S converted

waves (Rondenay, 2009 and references therein), is introduced to convert the precursory arrivals

of SS waves to the corresponding reflection depth and location (Gu et al., 2008; Heit et al.,

2010). The SS waveforms after the corrections for crust thickness and surface topography

correspond to equalized reflection at the Earth’s surface. Hence, each time sample preceding the

reference SS time can be mapped to a crustal/mantle depth according to the predicted travel-time

tables computed based on PREM (Dziewonski and Anderson, 1981) (Fig. 1B). The sampling

rate along the depth axis is 1 km.

Finally, to obtain a 3D reflectivity image we divide the study region into uniform, rectangular

Common Mid Point (CMP) gathers with horizontal and vertical step sizes of 2° and 8°,

respectively (IS THIS TRUE, AHMET?)_. Time-to-depth migration (Zheng et al., 2007) is

1

2

3

4

5

6

7

8

9

10

11

12

13

14

15

16

17

18

19

20

21

22

23

Page 8: University of Alberta › ~ygu › publications › japan_pa… · Web viewTracking Slabs Beneath Northeastern Pacific Subduction Zones Yu Jeffrey Gu University of Alberta, Department

subsequently performed at each cell and the entire set of resulting migrated traces is interpolated

using a 3D, bi-linear interpolation method provided by MATLAB. Despite linear interpolation

used in each direction, the bi-linear approach constructs new data points from a discrete set of

original data values based on a quadratic function (Press et al., 1993). The resolution of this

approach is further examined in the sections below.

3. Results

3.1. Maps of Reflection Amplitudes

Fig. 2 shows the region of interest in this study. Approximately 5000 high-quality traces are

retained after the data selection procedure detailed in the previous section. The ray theoretical

reflection points of the precursors (see Fig. 2) provide adequate resolution for the entire study

area. Furthermore, the increased data coverage in the latitude range of 35°-50° facilitates a

direct comparison of the mantle reflectivity structures in the vicinity of southern/central Japan

(cross-sections A and B) with those beneath the Kuril trench (cross-section C).

The Amplitude variations of 3D depth-converted SdS waves indicate the presence of large-scale

structures in the MTZ and shallow lower mantle. The top of the MTZ (Fig. 3) contains an

elongated, highly reflective zone (HRZ), extending from the northern Great Khingan Range in

the east to the northwestern corner of the study region beneath the Gobi desert. This 1500-km

wide anomaly reaches its maximum amplitude (9% of that of SS, for short, 9%) at ~425-km

depth, which is approximately 15 km below the global average of the 410-km seismic

discontinuity (Fig. 3A) (Gu et al., 2003; Houser et al., 2008). A second, weaker HRZs is visible

east of the Wadati-Benioff zone along the Kurile and Japan arcs, peaking at ~8% amplitude near

1

2

3

4

5

6

7

8

9

10

11

12

13

14

15

16

17

18

19

20

21

22

23

Page 9: University of Alberta › ~ygu › publications › japan_pa… · Web viewTracking Slabs Beneath Northeastern Pacific Subduction Zones Yu Jeffrey Gu University of Alberta, Department

the Hokkaido corner (see Fig. 3A).

The HRZs at the top of MTZ decays quickly with depth and the reflectivity pattern at ~520 km

depth is dominated by a strong (5-8%), uniquely shaped reflector (Fig. 3B). The center of this

reflector is located near Sikhote-Alin Mountains, roughly coinciding with the slab corner

between Japan and Kuril subduction zones outlined by Sam Gudmundsson and Sambridge

(1998) west of the Hokkaido corner (see depth map at 540 km, Fig. 3B). The orientation of this

boomerang-shaped structure (see map at 520 km) changes from ~30 deg oblique to the trench-

perpendicular direction west of Honshu Island to trench-perpendicular beneath northeastern

China. The vertical dimension of this mid-MTZ HRZ is no greater than 40 km (see Fig. 3B).

Large-scale reflective structures are clearly visible at the base of the upper mantle (Fig. 3C) and

below (Fig. 3d). Major north-south oriented HRZs are observed at 675-km depth northwest of

the Japan-Kuril arc-arc interaction region and the eastern section of the Gobi desert, respectively

(see Fig. 3C). The maximum amplitudes of both anomalies exceed 10%. The depths of the

HRZs indicate local depressions of 20+ km on the 660 beneath northeastern China. The

geographical locations of these HRZs roughly overlap with those of two lower-mantle reflectors

detectable at 900-930 km depths. The stronger and slightly deeper of the two HRZs (see 6%

amplitude isosurface, Fig. 3D) lies beneath the slab corner between Japan and Kuril subduction

zones. This semi-linear reflective structure is approximately trench-perpendicular and spans the

entire Wadati-Benioff zone in this arc-arc interaction region.

3.2. Correlation between reflectivity and seismic velocity

1

2

3

4

5

6

7

8

9

10

11

12

13

14

15

16

17

18

19

20

21

22

23

Page 10: University of Alberta › ~ygu › publications › japan_pa… · Web viewTracking Slabs Beneath Northeastern Pacific Subduction Zones Yu Jeffrey Gu University of Alberta, Department

Detailed information on the temperature-dependent seismic velocity and impedance-driven

reflectivity structure is necessary to accurately characterize mantle structure and processes near

subduction zones. To explore wave amplitude vs. velocity relationship, we overlay reflectivity

depth cross-sections (Fig. 4; see Fig. 2 for reference) with high-resolution regional P velocities

reported by Obayashi et al. (2006). While the use of a regional S velocity model would be ideal,

key mantle heterogeneities in the study region are better resolved by the high-resolution P wave

tomography (see review by Fukao et al., 2009). Reflections within the depth ranges 120-150 km,

380-440 km and 630-700 km are consistently observed in all cross-sections despite substantial

lateral variations in depth and amplitude. The focus of this study is on the MTZ and lower

mantle where waveform complexities associated with SS sidelobes are minimal (e.g., Shearer,

1993; Gu et al., 2003).

Fig. 4A shows highly undulating MTZ boundaries between the Pacific Plate and the volcanic arc

near central Honshu Island. The 410 east of the Japan trench undergoes 15-20 km local

depression relative to the cross-sectional average depth of 415 km.?? This 500-km wide HRZ

reaches the maximum reflection amplitude of ~8% beneath central Honshu Island, approximately

overlapping with a P wave low-velocity zone centered between 380-400 km depths (Obayashi et

al., 2006; see also Zhao and Ohtani, 2009; Li and van der Hilst, 2010; Bagly et al., 2009). The

reflectivity structure changes sharply toward the Wadati-Benioff zone where the 410 reaches

local minima in both depth (~395 km) and reflection amplitude (~5%) (see Fig. 4, Profile A).

Complex reflective structures are also evident at the base of the MTZ east of the Japan trench.

The 660 shows 25+ km peak-to-peak topography and the undulations appear to negatively

correlate with those of the 410 along the trench dip. Major depressions are identified beneath

1

2

3

4

5

6

7

8

9

10

11

12

13

14

15

16

17

18

19

20

21

22

23

Page 11: University of Alberta › ~ygu › publications › japan_pa… · Web viewTracking Slabs Beneath Northeastern Pacific Subduction Zones Yu Jeffrey Gu University of Alberta, Department

eastern Sea of Japan (~680 km) and Gulf of Chihii (~673 km) (see Figs. 3C and 4, Profile A),

with the former showing a slight offset from the center of predicted MTZ high velocities.

The shape of the high-velocity structure becomes quasi-linear near northern Honshu Island

where a significant number of deep-focus earthquakes have been recorded (Fig. 4, Profile B).

The 410 remains depressed in the east of the Wadati-Benioff zone (see profile A). A strong HRZ

is visible at ~300 km depth in this region, approximately outlining with the top of the low-

velocity zone (also see Fig. 3A) above the MTZ. The reflection characteristics of the 410 are

generally consistent with those from profile A, but the lateral variations in amplitude and depth

are visibly diminished relative to the former profile. At the base of the MTZ, the 660 shows

extreme local topography in the vicinity of the Wadati-Benioff zone. The depth of the 660

beneath the island arcs is ~645 km, the shallowest level in the entire profile, which significantly

reduces the MTZ width (~225 km) along the trench dip (see Fig. 4, Profile B). This anomalous

topographic structure on the 660 is accompanied by a broad depression beneath the Sea of Japan

and Changbai hotspot. The 1000-km wide structure west of the Hokkaido corner overlaps with a

P wave high-velocity zone near the base of the MTZ, but its lateral dimension is considerably

greater than that inferred from the 1+% P velocity variations.

The high-velocity structure beneath the Kuril subduction zone (Profile C) is visibly more

complex than those beneath the Japan subduction system, providing convincing evidence for 1) a

fast zone along slab dip that extends down to 750+ km depths, and 2) a horizontal MTZ anomaly

west of the Sea of Okhotsk with a possible ‘necking’ beneath the Sikhote-Alin Mountains. The

reflectivity structure in Profile C accentuates the complex slab morphology and kinematics in

1

2

3

4

5

6

7

8

9

10

11

12

13

14

15

16

17

18

19

20

21

22

23

Page 12: University of Alberta › ~ygu › publications › japan_pa… · Web viewTracking Slabs Beneath Northeastern Pacific Subduction Zones Yu Jeffrey Gu University of Alberta, Department

this region. Apparent reflection gaps are observed on the prodominantly continuous 410 and 660

along the Wadati-Benioff zone, with the latter anomaly nearly spanning the entire Sea of

Okhotsk. The shape of the 660 phase boundary west of this low-amplitude region closely

matches the outline of the 1% high-velocity structure in the MTZ (see Fig. 4, Profile C). We also

identify a highly undulating, piece-wise continuous lower mantle reflector beneath this profile,

showing the largest amplitude (~10%) beneath the reflection gap on the 660. The presence of

this lower-mantle reflector and isolated MTZ HRZs will be discussed in detail in Section 4.

The cross-sections shown by Fig. 4 (Profiles A-C) paint markedly different pictures of MTZ

reflectivity structures between Japan and Kuril subduction zones. A north-south transect over the

deepest part of the Wadati-Benioff zones (Fig. 4, Profile D) highlight the key observations that

differentiate between these two subduction systems. South of Hokkaido corner, large-scale

high-velocity structures appear to reside within the MTZ. Despite slightly reduced amplitudes,

the MTZ phase boundaries are generally detected and laterally continuous. In particular, the 660

is generally deeper than regional averages and the largest ‘visible’ depressions is detected

between the Korea Strait and Sea of Japan. On the other hand, the Kuril subduction zone

embodies a vertically continuous high-velocity structure that extends into the shallow lower

mantle. This P velocity anomaly is supported by a strong HRZ at ~930 km depth. Furthermore,

the amplitudes of the MTZ phase boundaries in the same regions are clearly below the threshold

of detection using SS precursors. It is worth noting that 1+% P velocities appear to reach a

depth of ~280 km, which coincides with a strong, possibly deformed, shallow mantle reflector

between the two subduction zones.

1

2

3

4

5

6

7

8

9

10

11

12

13

14

15

16

17

18

19

20

21

22

23

Page 13: University of Alberta › ~ygu › publications › japan_pa… · Web viewTracking Slabs Beneath Northeastern Pacific Subduction Zones Yu Jeffrey Gu University of Alberta, Department

A common observation between the Japan and Kuril subduction zone is the presence of mid

MTZ reflector(s) (see Fig. 4, Profiles A-D). We identify a single HRZ with maximum

amplitudes in excess of 6% at ~525 km near or within the Benioff zones in the southern profiles.

Two isolated mid-MTZ reflectors are present under the Kuril subduction zone at approximate

depth ranges of 500-530 km and 580-600 km, respectively. The depths of these reflectors vary

considerably in each profile, whereas the amplitudes generally increase from South to North.

3.3. Hypothesis testing and nominal resolution

Several procedures are implemented to ensure the stability and accuracy of the migration method

as well as the resolution of the SS precursor data set. To investigate the effect of earthquake

source and the migration algorithm, we compute synthetic seismograms (Fuchs and Muller,

1971; Kind, 1978; Hermann and Wang, 1985) for all source-station pairs based on PREM

(Dziewonski and Anderson, 1981) and earthquake source information from GCMT (see also

Section 2). The synthetic data set is then subjected to the same filtering, binning and migration

procedures as the actual observations. Fig. 5 shows the sample output for Profile C, which

validates at least two key premises of this study. First, the two bounding MTZ reflectors are

migrated to 400 and 670 km, respectively, to at least two decimal places. These values are

consistent with those of PREM, the 1D model used in the migration procedure, which suggest

that the time-to-depth mapping of the actual data is precise in the absence of lateral variations in

velocity or phase boundary topography. Furthermore, the amplitudes and depths of the MTZ

phase boundaries are nearly constant along the profile, which imply that the collective influence

of earthquake source mechanism, station response, and phase equalization on the results of data

migration stacks is negligible along this (see Figs. 5) and other (not shown) profiles.

1

2

3

4

5

6

7

8

9

10

11

12

13

14

15

16

17

18

19

20

21

22

23

Page 14: University of Alberta › ~ygu › publications › japan_pa… · Web viewTracking Slabs Beneath Northeastern Pacific Subduction Zones Yu Jeffrey Gu University of Alberta, Department

Questions have surfaced in recent years regarding the accuracy of the structure/topography

inferred from SS precursors due to the mini-max nature of reflected waves and their wide Fresnel

zones at long periods (Neele et al., 1997; Chaljub and Tarantola, 1997). Shearer et al. (1999)

addressed some of the potential biases through a multi-scale resolution analysis. By inverting for

synthetic differential travel times, they showed that a topographic inversion using long-period SS

precursor observations is virtually immune to smaller-scale artifacts at a major subduction zone.

Recent high-resolution images from the investigations of subduction slabs (Schmerr and

Garnero, 2007; Heit et al., 2010), hot mantle plumes (Schmrr and Garnero, 2006; Gu et al., 2009;

Cao et al., 2011) and lithosphere (Rychert and Shearer, 2009) are further testimonies of the

strong resolvability of SS precursors on finer-than-expected structures at mantle depths. Shear

waves has been known to resolve structures with length scales beyond their ‘nominal’ resolution,

especially when waveform information is incorporated (Ji and Nataf, 1998; Mégnin and

Romanowicz, 2000). In the case of SS precursors, minor errors are expected when relatively

large Fresnel zones of SS precursors collapse onto the fine grid adopted by this study, though the

lateral depth/amplitude differences between the averaging centers could persist and the apparent

connections between reflection amplitude, seismic velocity and seismicity (see Sections 3.1 and

3.2) are hard to dismiss as random occurrences.

Without repeating the successful experiment performed by Shearer et al. (1999), we examine

different CMP sizes to determine the optimal level of tradeoff between stability and resolution.

Fig. 6 shows the a comparison of reflectivity maps at 680 km based on averaging bins sizes of 2

x 6 = 12 deg2 (Fig. 6A) and 5 x 10 = 50 deg2 (Fig. 6B). Differences in the suggested spatial

1

2

3

4

5

6

7

8

9

10

11

12

13

14

15

16

17

18

19

20

21

22

23

Page 15: University of Alberta › ~ygu › publications › japan_pa… · Web viewTracking Slabs Beneath Northeastern Pacific Subduction Zones Yu Jeffrey Gu University of Alberta, Department

scales of the anomalies are apparent. A significant number of reflectors, some poorly resolved

due to insufficient data, exist in the former map whereas larger bin sizes tends to over-damp the

lateral variations in 660 topography. However, the location and maximum amplitudes of major

HRZ, e.g., a semi-linear structure across northern Honshu Island and a large, uniquely shaped

zone contouring the deepest part of the arc-arc interaction region, are minimally affected by bin

sizes. Our final choice of averaging area (32 deg2) represents an effective, albeit subjective,

compromise between image stability and resolution.

3.4. Uncertainty of reflectivity structure

We estimate the uncertainty of the reflectivity profiles based on bootstrapping resampling

algorithm (Efron, 1977). For each averaging bin, we first construct a ‘bootstrapped’ data set of

equivalent size to the original data set through random drawing. This procedure is performed

with the aid of a random generator (Press et al., 1992) and allows for repeated selections of the

same seismogram. We then perform data stacking and migration on this simulated data set and

obtain a single summary migrated seismogram for this particular averaging bin. This random

drawing and migration/averaging procedure is repeated 300 times in the same data gather to

obtain a statistically significant distribution of reflectivity at each depth. We estimate the

effective uncertainty by the standard deviation of these 300 bootstrapped seismograms (Efron,

1977; see also Shearer, 1993; Deuss and Woodhouse, 2002; Gu et al., 2003; Lawrence and

Shearer, 2006; An et al., 2007; Zheng et al., 2007), and apply the same treatment to all averaging

bins along each profile.

1

2

3

4

5

6

7

8

9

10

11

12

13

14

15

16

17

18

19

20

21

22

Page 16: University of Alberta › ~ygu › publications › japan_pa… · Web viewTracking Slabs Beneath Northeastern Pacific Subduction Zones Yu Jeffrey Gu University of Alberta, Department

The bootstrapped reflectivity profiles, which are constructed based on the average of the re-

sampled seismograms at each data gather, are nearly identical to the respective profiles shown by

Fig. 4. The bootstrapped uncertainties based on one standard deviation (Fig. 7) are generally

lower than 3% below 150 km. The spatial variation in uncertainty is nearly random, which

implies that the main MTZ reflectivity structures are reasonably well resolved in all profiles.

However, all four profiles show a 200-500 km wide section of increased uncertainties (reaching

~3% amplitude) that intercepts the seismogenic zone, e.g., beneath the Japan trench in Profiles A

and B and Strait of Tartary in Profile D (see Fig. 7). This anomalous zone is partly caused by

relatively sparse data coverage (see Fig. 2), though the scattering associated with inclined high-

velocity slab structures cannot be ignored. Further discussions of the latter effect will be

provided in Section 4.

4. Interpretation and discussion

Using reflected/scattered waves to illuminate the shape of major thermal and/or compositional

anomalies is analogous to ‘motion tracking’ in animation cinematography. In a nutshell, both

procedures take advantage of the relationships between reflection/scattering strengths and

changes in material properties including density, bulk or shear modulus and, in the case of

motion tracking, index of refraction of electromagnetic waves. Signals analyzed by both

applications are strongly sensitive to gradients in material properties and, with proper treatments,

can be used to decipher the shape or movements of the enclosed mass. On the other hand,

destructive interference or scattering of the waves caused by structural asperities could present

challenges, albeit providing additional information, to both applications. The incorporation of

additional physical constraints could be highly beneficial. For the case reflectivity imaging, the

1

2

3

4

5

6

7

8

9

10

11

12

13

14

15

16

17

18

19

20

21

22

23

Page 17: University of Alberta › ~ygu › publications › japan_pa… · Web viewTracking Slabs Beneath Northeastern Pacific Subduction Zones Yu Jeffrey Gu University of Alberta, Department

combination of reflectivity imaging and seismic tomography can substantially improve our

existing knowledge on morphology of subducted crust and lithosphere in the northwestern

Pacific region.

Many important factors must be considered in the discussion of the morphology and kinematics

of subducting slabs. From a mineralogical viewpoint, the slab geometry and the width of the

MTZ are strongly influenced by mineralogical phase transformations of olivine to wadsleyite

(near 400 km), wadsleyite to ringwoodite (near 520 km), and ringwoodite to perovskite +

magnesiowustite (near 660 km) (Katsura and Ito, 1989; Ita and Stixrude, 1992; Helffrich, 2000;

Bina, 2003 and references therein; Akaogi et al., 2007). The endothermic phase change at the

base of the MTZ increases local buoyancy forces, which can deflect subducting slabs and aid its

stagnation within the upper mantle (Christensen, 1995; Billen, 2008, 2010; Fukao et al. 2009).

Under thermodynamic equilibrium, a cold, water-rich slab is expected to raise the 410, depress

the 660 (due to the opposite signs of their Clapeyron slopes), and be responsible for a wide range

of reflective bodies within the mantle. The presence of water can strongly impact the phase

changes in the MTZ (e.g. Inoue et al. 1995; Kohlstedt et al., 1996; van der Meijde, 2003; Ohtani

et al. 2004; Kombayashi and Omori, 2006; Huang et al., 2006; Litasov et al., 2006; Suetsugu et

al., 2006). Below is a detailed account of some of the observed reflectivity structures in the

general framework of MTZ mineralogy and temperature.

4.1. Amplitudes of the MTZ discontinuities

The amplitudes of the reflections from the MTZ phase boundaries are functions of the impedance

contrast across the reflecting surface and the transition width. Furthermore, due to the

1

2

3

4

5

6

7

8

9

10

11

12

13

14

15

16

17

18

19

20

21

22

23

Page 18: University of Alberta › ~ygu › publications › japan_pa… · Web viewTracking Slabs Beneath Northeastern Pacific Subduction Zones Yu Jeffrey Gu University of Alberta, Department

summation of multiple seismograms at each location and the use of SS amplitude as the

normalization term, the topography on the interface and regional variations of SS can also

significantly impact the relative SS precursor amplitudes. This study exclusively focuses on the

positive reflections associated with increased material impedances with depth. This is a

subjective decision prompted by the simple observation that the signs of well-resolved

reflectivity structures are predominantly positive in our study area. Admittedly, many positive

phases are accompanied by sizeable negative peaks that could result from reductions in velocity

and/or density, e.g., near the top of a low velocity zone or the bottom of a high velocity structure.

We defer discussions of negative phases to a future study.

The detectable ranges of amplitudes are 4-9% for S410S and 4-12% for S660S, both showing

significant lateral variations. The former range overlaps with the predicted values of ~8% from

PREM (Dziewonski and Anderson) and global average of 6.7% (Shearer, 1996) based on SS

precursor observations, whereas the latter range falls well short of the predicted 14% (Shearer,

2000). These individual amplitude estimates are strongly affected by the strength of SS, the

normalizing reference phase. For instance, the presence of attenuating low-velocity structures

(e.g. Zhao et al., 1992, 1997, 2004; Lei and Zhao, 2005; Huang and Zhao, 2006), especially near

back arc regions (e.g., Xu and Wiens, 1997; Roth et al., 1999, 2000), could reduce the absolute

amplitude of SS and increase the relative amplitude. Compositional variations associated with

Al at the base of upper mantle (e.g., Weidner and Wang, 1997, 2000; Deuss and Woodhouse,

2002; Deuss, 2009) or Fe content (Akaogi et al., 2007; Inoue et al., 2010) are also known to

broaden phase boundary widths and cause reductions in precursor amplitudes. A more stable

parameter is the amplitude ratio between the 410 and the 660 (e.g., Shearer, 2000), which we

1

2

3

4

5

6

7

8

9

10

11

12

13

14

15

16

17

18

19

20

21

22

23

Page 19: University of Alberta › ~ygu › publications › japan_pa… · Web viewTracking Slabs Beneath Northeastern Pacific Subduction Zones Yu Jeffrey Gu University of Alberta, Department

estimate to be within the range of 0.7-0.8. This value is slightly higher than the earlier estimates

of 0.64-0.68 based on global SS precursor (shearer, 1996) and regional ScS observations

(Revenaugh and Jordan, 1991), but it is in poor agreement with that of PREM (0.5). A regionally

sharp 410 (e.g., Benz and Vidale, 1993; Vidale et al., 1995; Neele, 1996; Melbourn and

Helmberger, 1998; Ai and Zheng, 2003; Jasbinsek et al., 2010) could , although the presence of a

fluid-rich lens near the 410 (Smyth and Frost, 2002; van der Meijde, 2003; Inoue et al., 2010).

While these effects are difficult to constrain reliably based on seismic observations, scattering

associated with undulations on the two MTZ bounding discontinuities are more readily

observable (Shearer, 2000). The presence of dipping structures, particularly in the vicinity of

slabs, can preferentially lower the ‘perceived amplitude’ of the 660, hence the amplitude ratio of

410 vs. 660, due to the 25-30% larger topography on the 660 relative to that on the 410 (see Figs.

3 and 4). The following sections carefully examine discontinuity depths and their implications

for slab geometry and dyanmics.

4.2 Depth correlation of the MTZ discontinuities

The migrated reflectivity profiles provide new insights on the effect of mantle temperatures on

phase boundary variations. Results from high-pressure mineral physics (e.g., Katsura and Ito,

1989; Ita and Stixrude, 1992; Irifune et al., 1998; Helffrich, 2000; Akaogi et al., 2007) have

predicted a negative correlation, hence an increased transition width, between the phase

boundary undulations in an olivine-dominated mantle. Seismic evidence from regional (e.g., Li

et al., 2000; Collier et al., 2001; Lebedev et al., 2002; Saita et al., 2002; Ai et al., 2003; van der

Meijde et al., 2005; Ramesh et al, 2005; Tonegawa et al. 2005) and global (Shearer and Masters,

1992; Shearer, 1993; Gossler and Kind, 1996; Gu et al., 1998; Flanagan and Shearer, 1998;

1

2

3

4

5

6

7

8

9

10

11

12

13

14

15

16

17

18

19

20

21

22

23

Page 20: University of Alberta › ~ygu › publications › japan_pa… · Web viewTracking Slabs Beneath Northeastern Pacific Subduction Zones Yu Jeffrey Gu University of Alberta, Department

Lawrence and Shearer, 2006; Houser et al., 2008) analyses have generally supported this

hypothesis, but analyses based on lower-resolution approaches have largely attributed the

increased thickness to a strongly deformed 660 that correlates with the thermal variations at the

base of the upper mantle (Flanagan and Shearer, 1998; Gu et al., 1998, 2003; Gu and

Dziewonski, 2002; House et al., 2008). The depth of the 410 remains problematic in view of

mantle chemistry on all scales (e.g., Gilbert et al. 2002; Fee & Dueker 2004; Du et al. 2006; Gu

and Dziewonski, 2002; Gu et al., 2003; Deuss, 2007; Schmerr and Garnero, 2007; Tauzin et al.,

2008). Additional assumptions involving corrections (Flanagan and Shearer, 1998; Gu et al.,

2003; Schmerr and Garnero, 2006; Deuss, 2007; Houser et al., 2008) and/or mechanisms

predicated on extensive compositional variations (Schmerr and Garnero, 2007; Deuss, 2007; Gu

et al., 2009; Houser and Williams, 2010) are needed to reduce the difference between observed

and expected MTZ phase boundary perturbations.

To examine the correlation between temperature and discontinuity topography in our study area,

we focus on Profile A where both the 410 and 660 show the largest detectable topographic

variations and amplitudes near the Wadati-Benioffz zone (Fig. 8). The respective peak-to-peak

depth variations of the 410 and 660 are approximately 30 km and 410 km, which are comparable

to the largest variations reported by earlier global studies (Shearer, 1993; Gossler and Kind,

1997; Flanagan and Shearer, 1998; Gu et al., 2001, 2003; Houser et al., 2008; Lawrence and

Shearer, 2008). Both phase boundaries undergo extreme deformation from the trench onset to

the deepest part of the Wadati-Benioff zone across southern Japan (see Fig. 8A). A simple bin-

by-bin correlation assuming vertical thermal structures, the same approach used in the

aforementioned global studies, suggests a positive correlation between discontinuity depths over

1

2

3

4

5

6

7

8

9

10

11

12

13

14

15

16

17

18

19

20

21

22

23

Page 21: University of Alberta › ~ygu › publications › japan_pa… · Web viewTracking Slabs Beneath Northeastern Pacific Subduction Zones Yu Jeffrey Gu University of Alberta, Department

the length of the profile (see Fig. 4). To account for non-vertical structures following the slab dip

(~30 deg, Gudmundsson and Sambridge, 1998), we revise the correlation analysis by applying an

indexing change such that the depth of the 410 at a given location is correlated with the 660

depth at a location ~200 km further inland. The dip-corrected phase boundaries show clear

negative correlation in the vicinity of the slab (see Fig. 8A) and the corrected correlation

coefficient is -0.4 for the entire profile, a statistically significant value that clearly favors a

thermal origin for the observed MTZ topography. A key reason for the strong negative

correlation is the observed elevation of the 410 within the Wadati-Benioff zone. This feature

represents a major departure from those of earlier time-domain global studies of SS precursors

(e.g., Flanagan and Shearer, 1998; Gu et al. 2003), which we attribute to improved data

resolution in this study. From a broader perspective, this experiment not only highlights the

ability of SS precursors in resolving small-scale subduction zone anomalies, but also provides a

blueprint for to improve global correlation analyses via a priori information such as slab dip

angles.

4.3. Continuity of the 410 beneath northeast China

There have considerable discussion of results obtained from laboratory experiments on the

existence and support for a water/melt rich layer near the top of the MTZ (Wood, 1995; Inoue et

al., 1995, 2010; Kohlstedt et al., 1995; Smyth and Frost, 2002; Frost and Dolejs, 2007). Based on

these studies, wadsleyite has a strong capacity to accommodate hydroxyl (OH−), storing up to 3

wt.% H2O under equilibrium conditions (Wood, 1995; Inoue et al., 1995, 2010; Smyth and

Dolejs, 2007). These laboratory-based measurements have been supported by regional (e.g.,

Revenaugh and Sipkin, 1994; Zheng et al., 2007; Schmerr and Garnero, 2007; Schaeffer and

1

2

3

4

5

6

7

8

9

10

11

12

13

14

15

16

17

18

19

20

21

22

23

Page 22: University of Alberta › ~ygu › publications › japan_pa… · Web viewTracking Slabs Beneath Northeastern Pacific Subduction Zones Yu Jeffrey Gu University of Alberta, Department

Bostock, 2010) and global (Tauzin et al., 2010) seismic observations of low-velocity zones at

similar depths that cannot be sufficiently explained by thermal variations. The infiltration of

hydrous melt is further constrained through geodynamical calculation and synthesis (e.g.,

Bercovici and Karato, 2003; Karato, 2006; Leahy and Bercovici, 2007, 2010).

Our migrated reflectivity structures provide further regional constraints on this hypothesized

hydrous layer above the 410. The 410 west of the Wadati-Benioff zone (Fig. 8B) is consistently

shallower than the regional average in this study. The largest topography is observed in the

southernmost cross-section, reaching a depth of ~400 km beneath Korea and northeastern China.

The two northern profiles B and C show modest highs of ~410 km in the topography of the 410

near the Changbai hotspot and Sikhote-Alin Mountains, respectively. The average amplitudes of

the 410 in all three profiles far exceed the regional average, despite visible falloffs in the middle

of the highlighted section in the latter two profiles (see Fig. 8B). These characteristics are

reminiscent of those reported beneath the Tonga subduction zone (Zheng et al., 2007) based on

migrations of precursors to both P and S depth phases. However our highlighted section shows

strong positive reflections, which is opposite to those reported near Tonga, and the perturbations

in depth (<15 km relative to 410 km) is weaker than those presented by the earlier study (>20

km). Our highlighted region (see Fig. 8B) is also farther away from the Wadati-Benioff zone

than the target area in Zheng et al. (2007), though metasometism involving slab-derived fluids

rising through the flattened part of slabs (see Fukao et al., 2009 for review) could potentially be

as extensive as that beneath slab wedge. In fact, intraplate volcanoes nears Changbai mountains

and Wudalianchi region (see also Fig. 8B, Profile B) have been closely linked to processes

similar to back-arc spreading of the Japan slab (Lei and Zhao, 2005; Huang and Zhao, 2006).

1

2

3

4

5

6

7

8

9

10

11

12

13

14

15

16

17

18

19

20

21

22

23

Page 23: University of Alberta › ~ygu › publications › japan_pa… · Web viewTracking Slabs Beneath Northeastern Pacific Subduction Zones Yu Jeffrey Gu University of Alberta, Department

Schmerr and Garnero (2007) present another intriguing comparison. Based on multiple cross-

sections in South America, this earlier study inferred a ‘melt lens’ based on evidence of delayed

and split/missing S410S reflections east of the Nasca-South America convergent zone. The

presence of highly anomalous underside reflections received further support from Contenti et al.

(submitted, 2011) based on the method presented in this study. However, the complexity of the

S410S signal from South America far exceeds that from northeastern China. Should a fluid-rich

layer be present atop the MTZ beneath our study region, its spatial scale, infiltration/storage

mechanism and/or chemistry are likely to be different from those near Tonga and South America

subduction systems.

4.4. Slab stagnation and distortion

Subducted ocean basins in the western Pacific region have been known to deflect to a near-

horizontal direction the MTZ for nearly two decades (Okino et al., 1989; van der Hilst et al.,

1991; Fukao et al., 1992, 1993). Since then, ample evidence of slab stagnation (Fukao et al.,

1993, 2001) in subduction zones worldwide has been provided by global and regional

tomographic images with improved accuracy and resolution (Fukao et al., 2001, 2009; Zhao and

Ohtani, 2009; Li and van der Hilst, 2010; Sugioka et al., 2010) and anomalous dip-angle

variations suggested by the distribution of intermediate-depth earthquakes (Chen et al., 2004).

The conditions and characteristics of stagnant lithosphere have been constrained further through

numerical calculations incorporating thermo-petrological buoyancy forces (Tetzlaff and Schmeling, 2000;

Bina et al., 2001; Bina and Kawakatsu, 2010), rheology (Billen and Hirth, 2007; Billen, 2008), and plate

history and rollback (Torii and Yoshioka, 2007; Christensen, 2010; Zhu et al., 2010).

1

2

3

4

5

6

7

8

9

10

11

12

13

14

15

16

17

18

19

20

21

22

23

Page 24: University of Alberta › ~ygu › publications › japan_pa… · Web viewTracking Slabs Beneath Northeastern Pacific Subduction Zones Yu Jeffrey Gu University of Alberta, Department

With the help of seismic velocities, the reflectivity information provided by our study can place

crucial constraints on slab deformation at the base of the MTZ and the shallow lower mantle. In

particular, the shape of the HRZs near the 660 provides useful measures for the geometry and

dimension of the stagnant slabs. The two southern profiles presented in Fig. 3C and Fig. 4A-C

consistently show two distinct zones of large-lateral scale depression (Fig. 9), 1) near the

piercing point of the slab at the base of upper mantle, and 2) in the second half of the stagnant

slab inferred from recent tomographic models (e.g., Huang and Zhao, 2006; Fukao et al., 2009).

The two depressive zones have nearly identical shapes, particularly in Profile B, and depth of the

660 between them ranges from 655 to 660 km in both cases. Profile A shows significantly larger

topography than Profile B near the slab piercing point. For an isochemical mantle, the maximum

depth of ~685 km would suggest a temperature increase of XX-XX deg C depending on the

selected Clapeyron slope (REF). The reduction in topography from south (Profile A) to north

(Profile B) along the island arcs is in general agreement recent studies based on receiver

functions (Niu et al., 2005) and postcursors to sScS (Yamada and Zhao, 2007). The reduced

horizontal gradient in the topography of the 660 beneath northern Honshu could be caused by a

‘soft’ slab (Li et al., 2008) under the influence of trench migration and rollback. However, Li et

al. (2008) detected little or no oceanward broadening of the 660 from high-resolution S to P

converted waves. This is inconsistent with the apparent shift between the high-velocity contours

and the onset of the depressive zones in the vicinity of the island arcs (see Fig. 4 and Fig. 10).

Resolution differences of the two data sets (SS precursors vs. receiver functions) may be a

contributing factor, still, 100-300 km horizontal broadening/ponding of the Pacific slab at the

base of MTZ in the oceanward direction remains a strong possibility.

1

2

3

4

5

6

7

8

9

10

11

12

13

14

15

16

17

18

19

20

21

22

23

24

Page 25: University of Alberta › ~ygu › publications › japan_pa… · Web viewTracking Slabs Beneath Northeastern Pacific Subduction Zones Yu Jeffrey Gu University of Alberta, Department

A dimensional analysis of slab geometry based on the topography on the 660 is informative but

requires subjective definitions. Assume the points of intersection at 670 km depth mark the

corners of the topographic structures, we estimate the horizontal dimensions of depressive zones

to be 350-450 km in Profile A and 500-600 km in Profile B. The respective topographic highs

between the depressions are estimated to be ~700 km and ~400 km. The total length beyond the

depressions near the slab piercing point is approximately 1050 km for Profiles A and 900 km for

Profile B. These values are reasonably consistent with the estimated length of 800-1000 km for

deflected slab bodies (Huang and Zhao, 2006; Fukao et al., 2010), especially if slight reductions

due to horizontal averaging are considered in our estimates. However, as suggested by Fig. 10

and the estimates above, the truly ‘flat’ part of the slab that depresses the 660 phase boundary is

most-likely less than 600 km in width.

The migration-based topography of the 660 (see Fig. 9) challenges the ‘flatness’ of stagnant

slabs. The observation of contention is the average or shallow 660 between the depressive

zones, particularly in Profile A, whereas broad, continuous depression zones have been reported

earlier though seismic tomography (see Fukao et al., 2009 for review) and reflection depth/MTZ

thickness imaging (e.g., Shearer and Masters, 1992; Flanagan and Shearer, 1998; Gu et al.,

1998, 2003; Lawrence and Shearer, 2006; Houser et al., 2008). Furthermore, the amplitude of

the 660 within this uplifted region is consistently higher than the regional averages, which is

consistent with the expected decrease of ringwoodite-perovskite+magnisiowustite phase loop

under high-than-average temperatures. The observed phase boundary behavior is plausible

based on recent geodynamical calculations of slab geometry that consider 1) trench retreat

(Christensen 1996; Tagawa et al. 2007; Zhu et al., 2010) or 2) temperature- and pressure-dependent

1

2

3

4

5

6

7

8

9

10

11

12

13

14

15

16

17

18

19

20

21

22

23

24

Page 26: University of Alberta › ~ygu › publications › japan_pa… · Web viewTracking Slabs Beneath Northeastern Pacific Subduction Zones Yu Jeffrey Gu University of Alberta, Department

viscosity (Karato and Wu 1993; see Fig. 12 of Fukao et al. 2010). These calculations infer distinct

zones of depression at the slab piercing and re-entry points, between which the 660 remains

largely unperturbed. The images provided by these models are consistent with our observations

in the MTZ, though the expected reflections from the horizontally oriented slab segment in the

shallow lower mantle (e.g., Fukao et al., 2009) are not clearly observed from our data set (see

Fig. 9).

Alternatively, the internal undulations within stagnated slab body could suggest vertical

deformation of slab interface in the MTZ. Part of the lateral variations may be related to

advection (Kellogg et al. 1999; Obayashi et al., 2006), where the ambient and relatively hot

mantle material got ‘trapped’ during the interaction between the tip of the downgoing slab and

viscous lower mantle. Trench migration and rollback history could play a major role, as the

current geometry of stagnant slab could reflect changes in slab dip over the course of 100+ Ma

(see Schmid et al., 2002 for the case of Farallon plate subduction). Finally, the presence of water

(e.g., Listov et al. 2002, 2006; Inuoe et al., 2010) and possible separation of oceanic crust from

the downgoing lithosphere (Irifune and Ringwood, 1995;van Keken et al., 1996; Hirose et al., 1999,

2005) could also contribute to strong gradients in the topography of the 660 within the ‘flat’ part of the

slab.

4.5. Slab penetration beneath Kuril subduction zone

The reflectivity structures add new insights into the long-standing debate about the depth of slab

in the Pacific northwest (van der Hilst et al., 1991; Fukao et al., 1992; van der Hilst et al., 1997;

Fukao et al., 2001, 2009). While the vertical extent of slabs and the general style of mantle

convection remain debated on the global scale, there is growing evidence of scattered and

1

2

3

4

5

6

7

8

9

10

11

12

13

14

15

16

17

18

19

20

21

22

23

24

25

Page 27: University of Alberta › ~ygu › publications › japan_pa… · Web viewTracking Slabs Beneath Northeastern Pacific Subduction Zones Yu Jeffrey Gu University of Alberta, Department

deformed slab material in the lower mantle (van der Hilst et al., 1997; Bijwaard et al., 1998;

Fukao et al., 2001, 2009; Obayashi et al., 2006; Courtier and Revenaugh, 2008; Li and van der

Hilst, 2010; Chang et al., 2010).

Among the various HRZs documented in this study, MTZ anomalies contained in Profiles C and

D provide strong evidence for penetrating slabs in the western Pacific region. The most visible

change in the reflectivity structures from central Honshu slab to southern Kuril slab is the

amplitude reduction of the 410 and 660, highlighted by the apparent reflection gaps in Profiles C

and D. These gaps coincide with the Wadati-Benioff zone of the Kuril slab and their lateral

dimensions reflect the increasing width of the high velocity structure from the top to the bottom

of the MTZ (see Fig. 10A). The origin(s) of these reflection gaps remain(s) debatable. Factors

that have considerable impact on the amplitudes of the MTZ reflectors (see also Section 4.1)

include Al, water and Fe contents and optics.

There are merits and significant caveats in attributing the observed reflection gaps to variations

in mantle chemistry (e.g., the first three factors listed above). Under proper mantle conditions,

an increase in Al content could broaden the depth range of garnet-to-perovskite transformation

and influence olivine and pyroxene normaltive proportions near the base of the upper mantle

(Gasparik, 1996; Weidner and Wang, 1998; 2000). In a low temperature regime, e.g.,

subduction zones examined in this study, majorite garnet (a Al bearing mineral group) can

transform to metastable ilminite that eventually transforms to Ca-perovskite (e.g., Weidner and

Wang, 1998). These phase transitions exhibit different phase boundary behaviors from the

olivine system and adversely impact the interpretation of discontinuity depths and amplitudes.

1

2

3

4

5

6

7

8

9

10

11

12

13

14

15

16

17

18

19

20

21

22

23

Page 28: University of Alberta › ~ygu › publications › japan_pa… · Web viewTracking Slabs Beneath Northeastern Pacific Subduction Zones Yu Jeffrey Gu University of Alberta, Department

The presence of Al-bearing Akimotoite could introduce further complexities, e.g., a high velocity

layer or a steep velocity gradient, to mid MTZ depths at low temperatures (Gasparik, 1996;

Wang et al., 2004). However, changes in Al content mainly impact mantle reflectivity structure

under mid-to-lower MTZ pressure-temperature conditions (e.g., Weidner and Wang, 2000; Wang

et al., 2004). The restrictive condition greatly weakens the role of Al in view of the unexplained

absence of the 410 within Kuril slab.

Water transported into the MTZ by the subducting slab could also modify the impedance

contrast, hence the visibility of a reflecting body (van der Meijde, 2003; Ichiki et al., 2006).

Aided by strong capacities of wadleyite and ringwoodite to retain water (Inoue et al. 1995, Kohlstedt

et al. 1996; see Fukao et al., 2009 for review), a hydrous MTZ can simultaneously affect the width

and depth of the 660 (Litasov et al., 2006; Akaogi et al., 2007; Inoue et al., 2010). However, the

effect of water on the phase phase loop of the olivine-Wadsleyite transition is rather complex and

relatively minor with1 wt% H2O (Inoue et al., 2010). The implication is that a large amount of

water must be present in the descending slab to diminish the amplitude of S410S below the

detection threshold. Unfortunately, recent seismic observations (Fukao et al., 2009; Bina and

Kawakatsu, 2010), particularly those based on a novel modeling strategy for MTZ water content

(Suetsugu et al., 2006, 2010), have largely inferred ‘dry’ (e.g., <0.5%, Suetsugu et al., 2010)

slabs in various parts of the Pacific rim. Mechanism(s) predicated on increased Fe content in

slabs are similarly flawed. While increasing the Fe number can substantially broaden the phase

loops of both olivine-wadsleyite and ringwoodite-perovskite+magnisiowustite transitions

(Litasov et al., 2006; Akaogi et al. 2007; Inoue et al., 2010), the observational support for the

enrichment of Fe in subduction zones is not well established.

1

2

3

4

5

6

7

8

9

10

11

12

13

14

15

16

17

18

19

20

21

22

23

Page 29: University of Alberta › ~ygu › publications › japan_pa… · Web viewTracking Slabs Beneath Northeastern Pacific Subduction Zones Yu Jeffrey Gu University of Alberta, Department

The observed reflectivity gaps are best explained by effects commonly observed in optics.

Similar to the scattering of light, the observed amplitudes of the underside SH-wave reflections

are strongly influenced by the geometry of the reflecting surface. A dipping structure or

interface generally causes defocusing or scattering that, depending on the size of the structure

relative to the wavelength of the incoming wave, can result in the destructive interference of the

reflected/scattered waves. Therefore, local topography on the two MTZ bounding phase

boundaries in response to thermal and/or compositional variations are expected to tradeoff with

reflection amplitude obtained through averaging. This effect was documented by Chaljub and

Tarantola (1997) based on results from finite-difference modeling of S660S amplitude in

response to local topography and higher-than-average velocities, though the conclusions of that

study has been a subject of considerable debate (e.g., Shearer et al., 1999). We hereby quantify

the relationship between topography and SS precursor amplitude based on simulations of stacked

SS precursors from a depressed zone assuming uniform (case 1, Fig. 10A) and more extreme

(case 2, Fig. 10A) spatial distributions of reflection points. Reflectivity synthetic seismograms

(Randall are computed for common explosive source recorded by a station at 130-deg epicentral

distance. This experiment is repeated for depth perturbations (positive for the 410 and negative

for the 660) ranging from 0 (unperturbed PREM model) to 40 km. The resulting stacked

waveforms of SS precursors show a steady decay with increasing vertical topography,

particularly for case 2 where the reflection-point distribution is sparse (Fig. 10A and 10B). For

both cases, the amplitude drops to 50% for undulations of 15-25 km on the 410 and 25-35 km on

the 660, which will be problematic during the detection of large topographic features. Between

the two phase boundaries, the influence of btopography is larger for the 410 than the 660 due to a

1

2

3

4

5

6

7

8

9

10

11

12

13

14

15

16

17

18

19

20

21

22

23

Page 30: University of Alberta › ~ygu › publications › japan_pa… · Web viewTracking Slabs Beneath Northeastern Pacific Subduction Zones Yu Jeffrey Gu University of Alberta, Department

smaller assumed velocity jump at the former interface (see Fig. 10B). The amplitude decay

could be more severe for sparsely populated data (see case 2 simulations, Fig. 10). Furthermore,

the presence of large topography can significantly modify the waveform characteristics of the

superimposed seismogram. The wave shape broadens within increasing topography and,

depending on the frequency, can split into separate low-amplitude arrivals reflecting the top and

bottom of the topographic structure, respectively (see Fig. 10).

An underpinning message from Fig. 10 is that the maximum depth of the 660 could be 700 km or

deeper in the Pacific northwest (e.g., Revenaugh and Jordon, 1989; Niu et al., 2005). Based on

the impedance contrasts suggested by PREM, the amplitudes of both phase boundaries could

easily fall below the detection threshold of ~4% during the migration procedure when the

topography exceeds 35 km for the 660 and 20 km for the 410. While this is the ‘worst case’

scenario that assumes the averaging bin size is equivalent to the surface area of the topographic

structure, it does provide a viable explanation for the missing 410 and 660 within the Kuril slab.

The waveform splitting phenomenon (see Fig. 10) also has significant implications for the

detection of double reflectors. For example, results from high-pressure mineral physics (e.g.,

vacher et al., 1998; Weidner and Wang, 1998, 2000; Akaogi et al., 2002) have provided solid

laboratory evidence for garnet-ilmenite-perovskite transition near the base of the upper mantle.

Within low-temperature slabs, these garnet-related transitions are expected to take place over 60-

100 km range in depth (Vacher et al., 1998; Akaogi et al., 2002) that are capable of generating

mild reflections in seismic waves. Observationally, the occurrences multiple reflectors have

been reported under different tectonic settings (e.g., Deuss and Woodhouse, 2002; Ai and Zheng,

2003; Tibi et al., 2007), but their presence beneath northwest Pacific have been questioned

1

2

3

4

5

6

7

8

9

10

11

12

13

14

15

16

17

18

19

20

21

22

23

Page 31: University of Alberta › ~ygu › publications › japan_pa… · Web viewTracking Slabs Beneath Northeastern Pacific Subduction Zones Yu Jeffrey Gu University of Alberta, Department

(Lebedev et al., 2002; Tonegawa et al., 2005; Niu et al., 2005). In this study, only Kuril slab

(Profiles C and D, Fig. 4) show strongly dipped, weak reflecting bodies centered at ~700- and

780-km depths along the slab dip. These minor reflectivity structures are barely detectable,

showing ~4% amplitude each. While it is tempting to link these secondary structures to multiple

phase transitions, our numerical experiment above also cautions that the waveform complexities

associate with steep topographic structures should be considered in the interpretations.

The presence of a high-amplitude lower mantle HRZ beneath Kuril slab (Fig. 11) provide

potentially crucial support for the vertical extension of Kuril slab beyond the 660. Phase

transitions of Ca-perovskite (Stixrude et al., 2007), metastable garnet (Kawakatsu and Niu, 1994;

Kubo et al., 2002), as well as transformations of dense hydrous magnesium silicates under lower-

mantle pressure-temperature conditions, have been suggested as the origins of a series lower-

mantle reflectors (Shieh et al., 1998; Ohtani, 2005; Richard et al., 2006 and references therein).

The association of lower-mantle reflectors with phase changes is partially supported by the local

maxima of reflection amplitude beneath the reflection gap on the 660. However, reflections

from a sub-horizontal lower-mantle HRZ in northeast China between 850-1000 km depths

present a potential counter argument. The existence of a chemical boundary (Wen and

Anderson, 1997), which would influence the convective flow of mantle, cannot be ruled out.

We interpret the presence of the lower mantle reflector as an integral part of an ‘avalanching’

slab (Tackley, 1993) based on the following observations: 1) slab gaps at the 410 and 660 that

imply substantial mass and heat flux, 2) correlated fast velocity structure that maintains a strong

amplitude to depths comparable to that of the lower mantle reflector, 3) the presence of a strong

1

2

3

4

5

6

7

8

9

10

11

12

13

14

15

16

17

18

19

20

21

22

23

Page 32: University of Alberta › ~ygu › publications › japan_pa… · Web viewTracking Slabs Beneath Northeastern Pacific Subduction Zones Yu Jeffrey Gu University of Alberta, Department

(if not the strongest) lower mantle reflector in the vicinity of the slab gap. These observations

are self-consistent and could result from the same process (i.e., slab penetration) under different

pressure-temperature conditions and, possibly, mantle chemistry. Since the lower mantle HRZ

resides directly below the 660 reflection gap (rather than along the slab dip), the responsible

velocity/density structure could have undergone retrograde motion during its descend into the

lower mantle. These observations collectively defines the large difference between Kuril and

Honshu slabs in terms of maximum vertical extension.

4.6. Other HRZs and potential inferences

Two additional anomalous reflectivity structures from the SS migration images could have

significant implications for the mantle structure, dynamics and/or mineralogy if confirmed.

First, we identify one (Japan subduction zone) or multiple (Kuril region) mid-MTZ HRZ(s) with

reflection amplitudes of 5-9% within the MTZ (see Fig. 3B and Fig. 4). With an exception of

one instance east from the slab (see Fig. 4, Profile A), these HRZs are consistently detected

within the slab contours suggested by Obayashi et al. (2006). Reflective structures near 520-km

depth have been documented nearly 3 decades ago in the Pacific northwest from travel time

observations (Fukao et al., 1977). It was later proposed to be a mild global seismic discontinuity

based on pioneering studies of SS precursors (Shearer, 1990, 1991). Bock (1994) explained this

reflector as a potential data processing artifact due to strong low-frequency side-lobes of S410S

and S660S phases, though more recent results based on reflected and converted body waves

(Gossler and Kind, 1996; Shearer, 1996; Flanagan and Shearer, 1998; Gu et al. 1998, Chevrot et

al., 1999; Deuss and Woodhouse, 2001; Gu et al., 2003; Lawrence and Shearer, 2006, Deuss,

2009) have favored an explanation that involves regionally variable, highly undulating reflective

1

2

3

4

5

6

7

8

9

10

11

12

13

14

15

16

17

18

19

20

21

22

23

Page 33: University of Alberta › ~ygu › publications › japan_pa… · Web viewTracking Slabs Beneath Northeastern Pacific Subduction Zones Yu Jeffrey Gu University of Alberta, Department

structure(s) in the MTZ. In terms of mineral physics, this interface has been attributed to

wadsleyite to ringwoodite (Helffrich, 2000, Bina, 2003) and/or garnet to Ca-perovskite (Ita and

Stixrude, 1992) phase transitions. In cold mantle regions such as subduction zones, these

transformations likely occur at different MTZ depths (Saikia et al., 2008) and produce multiple

reflectors (Deuss and Woodhouse, 2001; Deuss, 2009). This may be the case for the observed

HRZs within the Kuril slab. Alternatively, delayed meta-stable olivine phase transition (Sung

and Burns, 1976; Iidaka and Suetsugu, 1992; Jiang et al., 2008, Bina and Kawakatsu; 2010) and

the presence of water within slabs are also viable source of enhanced reflections in active plate

convergence zones. Ultimately, an accurate interpretation of the anomalous HRZs within the

MTZ is predicated upon a greater consensus on the mantle condition surrounding slabs, for

example, the water content. In view of the apparent north-to-south difference between Japan (a

single 520 reflector) and Kuril (multiple reflectors) subduction zones, a combination of these

mechanisms may be needed to properly explain our observations in the Pacific Northwest.

Lastly, a narrow MTZ and a series of strong HRZs east of the Benioff-zone (see Figs. 4) both

suggest low MTZ temperatures. This interpretation is supported by findings in recent studies of

ScS reverberations (Revenaugh and Sipkin, 1994; Bagley et al. 2009), seismic tomography

(Obayashi et al., 2006; Huang and Zhao, 2006; Zhao and Ohtani, 2009), and electrical

conductivity (Ichiki et al., 2006). Furthermore, the strong reflection from these structures (8-

12% of SS) may not be sufficiently explained by a thermal origin alone. Compositional

variations associated with a hot mantle plume, which was once active during the past 130 Ma,

could provide the additional source material necessary to accomodate some of the strong

1

2

3

4

5

6

7

8

9

10

11

12

13

14

15

16

17

18

19

20

21

22

Page 34: University of Alberta › ~ygu › publications › japan_pa… · Web viewTracking Slabs Beneath Northeastern Pacific Subduction Zones Yu Jeffrey Gu University of Alberta, Department

reflections detected in the depth range of 250- 700 km (see Figs. 3 and 4) (Obayashi et al. 2006;

Honda et al., 2007; Bagley et al. 2009; Li and van der Hilst, 2010).

Conclusions

The dynamic processes beneath northwestern Pacific are only a microcosm of those beneath

many subduction systems globally. For this reason, inferences based on our high-resolution

reflectivity images could be potentially applicable to other regions with similar tectonic settings.

Based on the spatial correlation between reflectivity and seismic velocity, we conclude that the

origins of the majority of highly reflective zones are thermal, instead of compositional, in nature.

The combined reflectivity and velocity information enables us to detect and interpret the

geometry and strengths of major mantle heterogeneities in the approximate depth range of 300-

1000 km.

shows clear signs of bending within the MTZ, but the center of the stagnant section of the slab

appears to be deformed or folded, as suggested by an average or shallow 660. The depths of the

two MTZ bounding olivine phase boundaries are negatively correlated if slab dip is considered.

We also identify strong seismic reflector(s) within the slab body within the MTZ through out the

The Honshu slab does not appear to extend below the transition zone. negative overall

correlation between the depths of the two major olivine phase boundaries. However, localized

topography on the 660 within the presumed stagnant part of the slab suggests significant vertical

deformation near the base of the upper mantle. A single reflector is identified at the depth range

of 500-540 km, which could be associated with by changes in T . which causes strong negative

correlations of the olivine phase boundary but and Kuril slabs In particular, our analysis

1

2

3

4

5

6

7

8

9

10

11

12

13

14

15

16

17

18

19

20

21

22

23

Page 35: University of Alberta › ~ygu › publications › japan_pa… · Web viewTracking Slabs Beneath Northeastern Pacific Subduction Zones Yu Jeffrey Gu University of Alberta, Department

demonstrated that ‘gaps’ in the reflection amplitudes of the 410 and 660 are potentially

interconnected with anomalous lower mantle reflectors. Major mass/heat fluxes, large

topography on the base of upper mantle, and lower-mantle thermal/composition variations would

be expected at these locations. Intermittent reflections within the MTZ offer additional

information on the geometries and dynamics of stagnant slabs. In other words, a self-consistent

model of mantle processes beneath subduction zones is tenable from the presence, strengths, and

depths of mantle reflectors and their spatial correlations with seismic velocities.

From a technical standpoint, the results presented in this study provide a glimpse of the future for

regional-scale analysis based on intermediate-period SS precursors. Increasingly diverse

applications in recent years (e.g., Schmerr and Garnero, 2006, 2007; Houser et al., 2008;

Lawrence and Shearer, 2008; Gu et al., 2009; Rychert and Shearer, 2009; Heit et al., 2010; Cao

et al., 2010; Houser and Williams, 2010) have underlined the remarkable resolving power of this

data set, one that was traditionally tapped as a ‘low resolution’ constraint on mantle structure.

This trend will likely continue in the foreseeable future, especially in view of the growing

number of global seismic networks and applications of array methods.

Acknowledgement

We sincerely thank Suzan van der Lee for her constructive scientific input to this study. We are

grateful to Peter Shearer for his patience and professionalism in handling this manuscript. This

study also benefited from the helpful comments and suggestions from Nicholas Schmerr and an

anonymous reviewer, as well as from the technical assistance from the IRIS Data Management

1

2

3

4

5

6

7

8

9

10

11

12

13

14

15

16

17

18

19

20

21

22

Page 36: University of Alberta › ~ygu › publications › japan_pa… · Web viewTracking Slabs Beneath Northeastern Pacific Subduction Zones Yu Jeffrey Gu University of Alberta, Department

Center. This project is jointly funded by CFI, Alberta Innovates, Alberta Geological Survey,

National Science and Engineering Council (NSERC), and the University of Alberta.

References

Ai, Y., Zheng, T., 2003. The upper mantle discontinuity structure beneath eastern China.

Geophysical Research Letters 30, doi:10.1029/2003GL017678.

Akaogi, M., A. Tanaka, and, E. Ito, Garnet-ilmenite-perovskite transitions in the systemMg4Si4O12-Mg3Al2Si3O12 at high pressures and high temperatures: phase equilibria, calorimetryand implications for the mantle structure, Phys. Earth Planet. Inter., 132, 303-324, 2002.

Akaogi, M., Takayama, H. Kojitani, H. Kawaji, H., Atake, T., 2007. Low-temperature heat

capacities, entropies and enthalpies of Mg2SiO4 polymorphs, and and post-spinel

phase relations at high pressure. Physics and Chemistry of Minerals 34, 169-183.

Bagly, B., 1982. Geometry of subducted plates and island arcs viewed as a buckling problem. Geology 10, 629-632.

Bagley, B., Courtier, A. M., Revenaugh, J., 2009. Melting in the deep upper mantle oceanward of

the Honshu slab. Physics of the Earth and Planetary Interiors 175, 137–144.

Bercovici, D., and S. Karato (2003), Whole mantle convection and thetransition‐zone water filter, Nature, 425, 39–44.

Bassin C., Laske, G., Masters, G., 2000. The current limits of resolution for surface wave

tomography in North America. EOS Trans AGU 81, F897.

H.M. Benz, J.E. Vidale, Sharpness of upper-mantle discontinuitiesdetermined from high-frequency reflections, Nature 365(1993) 147–150.

1

2

3

4

5

6

7

89

1011

12

13

14

15

161718

19

20

212223

24

25

26

27282930

31

Page 37: University of Alberta › ~ygu › publications › japan_pa… · Web viewTracking Slabs Beneath Northeastern Pacific Subduction Zones Yu Jeffrey Gu University of Alberta, Department

Bijwaard, H.,Spakman, W., Engdahl, E. R., 1998. Closing the gap between regional and global

travel time tomography. Journal of Geophysical Research 103, 30055–30078.

Billen, M. I., 2008. Modeling the dynamics of subducting slabs. Annual Review of Earth and

Planetary Sciences 36, 325-356.

Billen, M. I., Hirth, G., 2007. Rheologic controls on slab dynamics. Geophysics, Geochemistry,

Geosystems 8, Q08012, doi:10.1029/2007GC001597.

Bina, C.R., Stein, S., Marton, F.C., Van Ark, E.M., 2001. Implications of slab mineralogyfor subduction dynamics. Phys. Earth Planet. Inter. 127, 51–66.

Bina, C. R., 2003. Seismological Constraints upon Mantle Composition, (in) Carson R. (Ed),

Treatise on Geochemistry vol 2., Elsevier Science Publishing, Oxford, pp. 39-59.

Bina, C. R., Kawakatsu, H., 2010. Buoyancy, bending and seismic visibility in deep slab

stagnation. Physics of the Earth and Planetary Interiors, 183, 330-340.

Bird, P., 2003. An updated digital model of plate boundaries. Geochemistry, Geophysics,

Geosystems 4, 1027.

Cao, Q., Van der Hilst, R.D., De Hoop, M.V., Shim, S.-H., 2011. Seismic imaging of transition

zone discontinuities suggests hot mantle west of Hawaii. Science 332, 1068-1071,

doi:10.1126/science.1202731.

Chaljub, E., Tarantola, A., 1997. Sensitivity of SS precursors to topography on the upper-mantle 660-km discontinuity. Geophysical Research Letters 24, 2613-2616.

Chang, S-J., van der Lee, S., Flanagan, M. P., Bedle, H., Marone, F., Matzel, E. M., Pasyanos, M.

E., Rodgers, A. J., Romanowicz, B., Schmid, C., 2010. Joint inversion for 3-dimensional S-

velocity mantle structure along Tethyan margin. Journal of Geophysical Research 115, B08309.

Chen, P., Bina, C.R., Okal, E.A., 2004. A global survey of stress orientations in subductingslabs as revealed by intermediate-depth earthquakes. Geophys. J. Int. 159,

1

2

3

4

5

6

7

8

91011

12

13

14

15

16

17

18

19

20

21

22

23

24

25

262728

29

30

31

323334

Page 38: University of Alberta › ~ygu › publications › japan_pa… · Web viewTracking Slabs Beneath Northeastern Pacific Subduction Zones Yu Jeffrey Gu University of Alberta, Department

721–733.

Chen, L., and Y. Ai (2009), Discontinuity structure of the mantle transition zone

beneath the North China Craton from receiver function migration, J. Geophys.

Res., 114, B06307, doi:10.1029/2008JB006221.

Christensen, U., 1995. Effects of phase transitions on mantle convection. Annu. Rev. Earth

Planet. Sci. 23, 65–87.

Christensen UR. 1996. The influence of trench migration on slab penetration into the lower mantle. EarthPlanet. Sci. Lett. 140:27–39

Christensen, U., 2010. Geodynamic models of deep subduction. Phys. Earth Planet.Inter. 127, 25–34.

Collier, J., Helffrich, G., Wood, B., 2001. Seismic discontinuities in subduction zones.Phys. Earth Planet. Int. 127, 39–49.

Courtier, A. M., Revenaugh, J., 2006. A water-rich transition zone beneath the eastern United

States and Gulf of Mexico from multiple ScS reverberations. Earth’s Deep Water Cycle. Ed. S.D.

Jacobsen and S. van der Lee. Geophysical Monograph (AGU), 181–193.

Courtier, A. M., Revenaugh, J., 2008. Slabs and shear wave reflectors in the mid-mantle. Journal

of Geophysical Research 113, B08312.

DeMets, C., Gordon, R.G., Argus, D.F., Stein, S., 1990. Current plate motions. GeophysicalJournal International 101, 425–478.

Deuss, A. and Woodhouse, J. H., 2002. A systematic search for mantle discontinuities using SS-

precursors. Geophysical Research Letters, 29, 1249-1252.

1

2

3

4

5

6

7

8

91011

121314

15

161718

19

20

21

22

23

24

25

26

27

282930

31

32

33

34

Page 39: University of Alberta › ~ygu › publications › japan_pa… · Web viewTracking Slabs Beneath Northeastern Pacific Subduction Zones Yu Jeffrey Gu University of Alberta, Department

Deuss, A., 2007. Seismic observations of transition zone discontinuities beneath

hotspot locations, Special Papers of the Geological Society of America, 430

(Plates, plumes and Planetary Processes edited by Foulger, G. R. and D. M.

Jurdy), 121-136.

Deuss, A., 2009. Global observations of mantle discontinuities using SS and PP precursors.

Surveys in Geophysics, 30 (4-5), 301-326.

Du, Z., Vinnik, L.P. & Foulger, G.R., 2006. Evidence from P-to-S mantleconverted waves for a flat ‘660-km’ discontinuity beneath Iceland, Earthplanet. Sci. Lett., 241, 271–280.

Dziewonski, A. M., Anderson, D. L., 1981. Preliminary reference earth model, Physics of the

Earth and Planetary Interiors 25, 297-356.

Dziewonski, A. and Woodhouse, J. (1983). Studies of the seismic source using normal-mode

theory. In Kanamori, H. and Boschi, E., editors, Earthquakes: observation, theory, and

interpretation: notes from the International School of Physics ``Enrico Fermi'' (1982: Varenna,

Italy), pages 45-137. North-Holland Publ. Co., Amsterdam.

Efron, B., 1977. Bootstrap method, another look at the Jackknife. The Annals of Statistics 7, 1-

26.

Fee, D. & Dueker, K., 2004. Mantle transition zone topography and structurebeneath the Yellowstone hotspot, Geophys. Res. Lett., 31, L18603,doi:10.1029/2004GL020636.

Flanagan, M.P., Shearer, P. M., 1998. Global mapping of topography on transition zone velocity

discontinuities by stacking SS precursors. Journal of Geophysical Research 103, 2673–2692.

Frost, D. 2008. The upper mantle and transition zone. Elements 4, 171-176.

Frost, D., Dolejs, D., 2007. Experimental determination of the effect of H2O on the 410-km

seismic discontinuity. Earth and Planetary Science Letters, 256, 182-195.

1

2

3

4

5

6

7

89

1011

12

13

14

15

16

17

18

19

20

21

22

23242526

27

28

29

30

31

32

33

34

Page 40: University of Alberta › ~ygu › publications › japan_pa… · Web viewTracking Slabs Beneath Northeastern Pacific Subduction Zones Yu Jeffrey Gu University of Alberta, Department

Fuchs, K., and G. Muller, Computation of synthetic seismograms with the reflectivity method and comparison with observations, Geophys. J. R. Astr. Soc., 23, 417-433, 1971.

Fukao, Y., 1977. Upper mantle P structure on the ocean side of the Japan-Kuril Arc. Geophysical

Journal of Royal Astronomical Society 50, 621-642.

Fukao, Y., Obayashi, M., 2010. Transition from slab stagnation to penetration beneath the

northwestern Pacific and South America. American Geophysical Union, Fall Meeting, abstract

#DI23C-01.

Fukao, Y., Obayashi, M., Inoue, H., Nenbai, M., 1992. Subducting slabs stagnant in the Mantle

Transition Zone. Journal of Geophysical Research 97, 4809–4822.

Fukao, Y.,, Obayashi, M., Nakakuki, T., Deep Slab Project Group, 2009. Stagnant slab: A review.

Annual Review of Earth and Planetary Sciences 37, 19-46.

Fukao, Y., Widiyantoro, S., Obayashi, M., 2001. Stagnant slabs in the upper and lower mantle

transition region. Reviews of Geophysics 39, 291–323.

Gilbert, J.H., Sheehan, A.F., Dueker, K.G. & Molnar, P., 2003. Receiverfunctions in the western United States, with implications for uppermantle structure and dynamics, J. geophys. Res., 108(B5), 2229,doi:10.1029/2001JB001194.Gossler, J. & Kind, R., 1996.

Gasparik, T., 1996. Melting experiments on the enstatite-diopsidejoin at 70–224 kbar, including the melting of diopside. Contrib.Mineral. Petrol. 124, 139–153.

Gorbatov A, Kennett BLN. 2002. Joint bulk-sound and shear tomography for western Pacific subductionzones. Earth Planet. Sci. Lett. 210:527–43

Gorbatov A,Widiyantoro S, Fukao Y, Gordeev E. 2000. Signature of remnant slabs in the North Pacific fromP-wave tomography. Geophys. J. Int. 142:27–36

Gossler J., Kind, R., 1996. Seismic evidence for very deep roots of continents. Earth and

Planetary Science Letters 138, 1-13.

1

234

5

6

7

8

9

10

11

12

13

14

15

16

17

18

192021222324

25262728

293031

323334

35

36

37

Page 41: University of Alberta › ~ygu › publications › japan_pa… · Web viewTracking Slabs Beneath Northeastern Pacific Subduction Zones Yu Jeffrey Gu University of Alberta, Department

Gu Y. J., Dziewonski, A. M., Agee, C. B., 1998. Global de-correlation of the topography of

transition zone discontinuities. Earth and Planetary Science Letters 157, 57-67.

Gu Y. J., Dziewonski, A. M., 2002. Global variability of transition zone thickness. Journal of

Geophysical Research 107, 2135.

Gu, Y. J., Dziewonski, A. M., Ekström, G., 2003. Simultaneous inversion for mantle shear

velocity and topography of transition zone discontinuities. Geophysical Journal International

154, 559–583.

Gu, Y. J., Schultz, R., Okeler, A., 2008. Migration and radon imaging of the western Pacific

subduction zones using SdS waves. EOS Trans AGU. 89, DI12A-08.

Gudmundsson O., Sambridge, M., 1998. A regionalized upper mantle (RUM) seismic model.

Journal of Geophysical Research 103, 7121-7136.

Heit, B., Yuan, X., Bianchi, M., Kind, R. And Gossler, J., 2010. Study of the lithospheric and

upper-mantle discontinuities beneath eastern Asia by SS precursors. Geophysical Journal

International 183, 252-266.

Helffrich, G., 2000. Topography of the transition zone seismic discontinuities. Reviews of

Geophysics 38, 141–158.

Herrmann, R. B., and C. Y. Wang, A comparison of synthetic seismograms, Bull. Seism. Soc. Am., 75, 41-

56, 1985.

Hirose K, Fei Y, Ma Y, Mao HK. 1999. The fate of subducted basaltic crust in the Earth’s lower mantle.Nature 397:53–56

Hirose K, Takafuji N, Sata N, Ohishi Y. 2005. Phase transition and density of subducted MORB crust in thelower mantle. Earth Planet. Sci. Lett. 237:239–51

1

2

3

4

5

6

7

8

9

10

11

12

13

14

15

16

17

18

19

20

21

22

23

24

25

26

272829

303132

33

Page 42: University of Alberta › ~ygu › publications › japan_pa… · Web viewTracking Slabs Beneath Northeastern Pacific Subduction Zones Yu Jeffrey Gu University of Alberta, Department

Honda, S., Morishige, M., Orihashi, Y., 2007. Sinking hot anomaly trapped at the 410 km

discontinuity near the Honshu subduction zone, Japan. Earth and Planetary Science Letters 261,

565–577.

Houser, C., Williams, Q., 2010. Reconcilling Pacific 410 and 660 km discontinuity topography,

transition zone shear velocity patterns, and mantle phase transitions. Earth and Planetary Science

Letters 296, 255-266.

Houser, C., Masters, G., Flanagan, M., Shearer, P., 2008. Determination and analysis of long-

wavelength transition zone structure using SS precursors, Geophysical Journal International 174,

178-194.

Huang, J., Zhao, D., 2006. High-resolution mantle tomography of China and surrounding

regions. Journal of Geophysical Research 111, B09305.

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. Physics of The Earth and Planetary Interiors 155, 1 – 15.

Iidaka, T., Suetsugu, D., 1992. Seismological evidence for metastable olivine inside a subducting

slab. Nature 356, 593–595.

T. Inoue, H. Yurimoto, Y. Kudoh, Hydrous modified spinel,Mg1.75SiH0.5O4: a new water reservoir in the mantle transitionregion, Geophys. Res. Lett. 22 (1995) 117–120.

Inoue, T., Ueda, T., Tanimoto, Y., Yamada, A., Irifune, T., 2010. The effect of water on the high-

pressure phase boundaries in the system Mg2SiO4-Fe2SiO4. Journal of Physics: Conference Series

215, doi:10.1088/1742-6596/215/1/012101.

Irifune T, Ringwood AE. 1993. Phase transformations in subducted oceanic crust and buoyancy relationshipsat depths of 600–800 km in the mantle. Earth Planet. Sci. Lett. 117:101–10

1

2

3

4

5

6

7

8

9

10

11

12

13

14

15

16

17

18

19

20

21

22

23242526

27

28

29

30

313233

Page 43: University of Alberta › ~ygu › publications › japan_pa… · Web viewTracking Slabs Beneath Northeastern Pacific Subduction Zones Yu Jeffrey Gu University of Alberta, Department

Ita, J., Stixrude, S., 1992. Petrology, elasticity, and composition of the mantle transition zone.

Journal of Geophysical Research 97, 6849–6866.

Irifune, T., Nishiyama, N., Kuroda, K., Inoue, T., Isshiki, M., Utsumi, W., Funakoshi, K.,Urakawa, S., Uchica, T., Katsura, T., Ohtaka, O., 1998. The postspinel phase boundary inMg2SiO4 determined by in situ X-ray diffraction. Science 279, 1698–1700.

Ji, Y., Nataf, H. C. 1998. Detection of mantle plumes in the lower mantle by diffraction tomography: Hawaii. Earth and Planetary Science Letters 159, 99-115.

Jasbinsek, J. J., Dueker, K. G., Hansen, S. M., 2010. Characterizing the 410 km discontinuity

low-velocity layer beneath the LA RISTRA array in the North American Southwest.

Geochemistry, Geophysics, Geosystems 11, Q03008, doi:10.1029/2009GC002836.

Jiang, G., Zhao, D., Zhang, G., 2008. Seismic evidence for a metastable olivine wedge in the

subducting Pacific slab under Japan Sea. Earth and Planetary Science Letters 270, 300–307.

Jordan, T. H. (1977). Lithospheric slab penetration into the lower mantle beneath the Sea of Okhotsk. Journal of Geophysics - Zeitschrift Fur Geophysik. Vol. 43 (1-2), pp. 473-496.

Kaneshima S, Helffrich G. 1999. Dipping low-velocity layer in the midlower mantle: evidence for geochemicalheterogeneity. Science 283:1888–91.

S. Karato, Remote sensing of hydrogen in the Earth's mantle,Rev. Mineral. Geochem. 62 (2006) 343–375.

Karato S,Wu P. 1993. Rheology of the upper mantle: a synthesis. Science 260:771–78

Katsura, T., Ito, E., 1989. The system Mg2SiO4-Fe2SiO4 at high pressures and temperatures; precise determination of stabilities of olivine, modified spinel, and spinel. Journal of Geophysical Research 94, 15663-15670.

Kawakatsu, H., Niu, F., 1994. Seismic evidence for a 920-km discontinuity in the mantle. Nature

371, 301–305.

Kellogg, L. H., Hager, B. H., van der Hilst, R. D., 1999. Compositional stratification in the deep

mantle. Science 283, 1881-1884.

1

2

3

4567

8

9101112

13

14

15

16

17

18

19

202122

232425

262728

2930

31323334

35

36

37

38

Page 44: University of Alberta › ~ygu › publications › japan_pa… · Web viewTracking Slabs Beneath Northeastern Pacific Subduction Zones Yu Jeffrey Gu University of Alberta, Department

Kind, R., The reflectivity method for a buried source, J. Geophys., 44, 603-612, 1978.

D.L. Kohlstedt, H. Keppler, D.C. Rubie, Solubility of water inthe α, β and γ phases of (Mg,Fe)2SiO4, Contrib. Mineral. Petrol.123 (1996) 345–357.

Kubo, T., Ohtani, E., Kondo, T., Kato, T., Toma, M., Hosoya, T., Sano, A., Kikegawa, T., Nagase,

T., 2002. Metastable garnet in oceanic crust at the top of the lower mantle. Nature 420, 803–806.

Lawrence, J.F., and M. E. Wysession, Seismic evidence for subduction-transported water in the

lower mantle, in Earth's Deep-Water Cycle. AGU Monograph, 251-261, 2006.

Lawrence, J.F., and P.M. Shearer, A global study of transition zone Thickness using receiver

functions, J. Geophys. Ress., 111, dio:10.1029/2005JB003973, 2006.

Lawrence, J.F., and P.M. Shearer, Imaging mantle transition zone thickness with SdS-SS finite-

frequency sensitivity kernels, Geophys. J. Int., 174, 143-158, doi: 10,1111/j.1365-

246X.2007.03673.x, 2008.

Leahy, G. M., and D. Bercovici (2007), On the dynamics of a hydrous meltlayer above the transition zone, J. Geophys. Res., 112, B07401,doi:10.1029/2006JB004631.

Leahy, G. M., and D. Bercovici (2010), Reactive infiltration of hydrous melt above the mantle transition zone, J. Geophys. Res., 115, B08406,doi:10.1029/2009JB006757.

Lebedev, S., Chevrot, S., van der Hilst, R. D., 2006. The 660-km discontinuity with the

subducting NW-Pacific lithospheric slab. Earth and Planetary Science Letters 205, 25-35.

Lei, J., Zhao, D., 2005. P-wave tomography and origin of the Changbai intraplatevolcano in Northeast Asia. Tectonophysics 397, 281–295.

Li, C., van der Hilst, R. D., 2010. Structure of the upper mantle and transition zone beneath

Southeast Asia from traveltime tomography. Journal of Geophysical Research 115, B07308.

1

23456

7

8

9

10

11

12

13

14

15

16

17

18

19

20212223

24252627

28

29

30

3132

33

34

35

36

Page 45: University of Alberta › ~ygu › publications › japan_pa… · Web viewTracking Slabs Beneath Northeastern Pacific Subduction Zones Yu Jeffrey Gu University of Alberta, Department

Li, J., Chen, Q.-F., Vanacore, E, Niu, F., 2008. Topography of the 660-km discontinuity and

impliations for a retrograde motion. Geophysical Research Letters 35, L01302, doi:10.1029/

2007GL031658.

Li, X., Sobolev, S. V., Kind, R., Yuan, X., Estabrook, C. H., 2000. A detailed receiver function

image of the upper mantle discontinuities in the Japan subduction zone. Earth and Planetary

Science Letters 183, 527–541.

Litasov, K., and E. Ohtani (2002), Phase relations and melt compositions inCMAS‐pyrolite‐H2O system up to 25 GPa, Phys. Earth Planet. Inter.,134, 105–127.

Litasov KD, Ohtani E, Sano A. 2006. Influence of water on major phase transitions in the Earth’s mantle. SeeJacobsen & van der Lee 2006, pp. 95–111

Mégnin, C., Romanowicz, B., 2000. The 3D shear velocity structure of the mantle from the

inversion of body, surface and higher mode waveforms. Geophysical Journal International 143,

709-728.

Melbourne, T., Helmberger, D., 1998. Fine structure of the 410-km discontinuity. Journal of

Geophysical Research 103, 10091-10102.

F. Neele, Sharp 400-km discontinuity from short-period Preflections, Geophys. Res. Lett. 23 (1996) 419–422.

Neele, F., de Regt, H.,Van Decar, J., 1997. Gross errors in upper-mantle discontinuity topography from underside reflection data, Geophysical Journal International 129, 194-204.

Niu, F., Levander, A., Ham, S., Obayashi, M., 2005. Mapping the subducting Pacific slabbeneath southwest Japan with Hi-net receiver functions. Earth and Planetary Science Letters 239, 9–17.

Northrup, C. J., Royden, L. H., Burchfiel, B. C., 1995. Motion of the Pacific plate relative to

Eurasia and its potential relation to Cenozoic extension along the eastern margin of Eurasia.

Geology 23, 719–722.

1

2

3

4

5

6

7

89

1011

121314

15

16

17

18

19

20

21

222324

25

26272829303132

33

34

35

36

Page 46: University of Alberta › ~ygu › publications › japan_pa… · Web viewTracking Slabs Beneath Northeastern Pacific Subduction Zones Yu Jeffrey Gu University of Alberta, Department

Obayashi, M., Sugioka, S., Yoshimitsu, J., Fukao, Y., 2006. High temperature anomalies

oceanward of subducting slabs at the 410-km discontinuity. Earth and Planetary Science Letters

243, 149–158.

Ohtani, E., 2005. Water in the Mantle. Elements 1, 25–30.

Okino K, Ando M, Kaneshima S, Hirahara K. 1989. The horizontally lying slab. Geophys. Res. Lett. 16:1059–2062

Press, W. H., Flannery, B. P., Teukolsky, S. A., Vetterling, W. T.. 1992. Numerical recipes in C:

The Art of scientific computing. 2nd Edition, Cambridge University Press, Cambridge, United

Kingdom.

Ramesh, D., Kawakatsu, H., Watada, S., Yuan, X., 2005. Receiver function images of thecentral Chugoku region in the Japanese islands using Hi-net data. Earth PlanetsSpace 57, 271–280.

Revenaugh, J., Jordan, T. H., 1991. Mantle layering from ScS reverberations: 2. the transition

zone. Journal of Geophysical Research 96, 19763-19810.

Revenaugh, J., Sipkin, S. A., 1994. Seismic evidence for silicate melt atop the 410-km mantle

discontinuity. Nature 369, 474–476.

Richard, G., Bercovici, D., Karato, S -I., 2006. Slab dehydration in the Earth’s mantle transition

zone. Earth and Planetary Science Letters 251, 156–167.

Rondenay, S., 2009. Upper mantle imaging with array recordings of converted and scattered

teleseismic waves. Survey of Geophysics 30, 377.

Rychert, C. A., Shearer, P. M., 2009. A global view of the lithosphere-asthenosphere boundary.

Science 332, 495-498, doi:10.1126/science.1169754.

Saita, T., Suetsugu, D., Ohtaki, T., Takenaka, H., Kanjo, K., Purwana, I., 2002. Transitionzone thickness beneath Indonesia as inferred using the receiver function methodfor data from JISNET regional broadband seismic network. Geophys. Res. Lett. 29.doi:10.1029/2001GL013629.

1

2

3

4

5

678

9

10

11

12

13141516

17

18

19

20

21

22

23

24

25

26

27

28

29

30

31

3233343536

Page 47: University of Alberta › ~ygu › publications › japan_pa… · Web viewTracking Slabs Beneath Northeastern Pacific Subduction Zones Yu Jeffrey Gu University of Alberta, Department

Schaeffer, A. J., Bostock, M. G., 2010. A low velocity zone atop the transition zone in

northwestern Canada. Journal of Geophysical Research 15, B06302,

doi:10.1029/2009JB006856.

Schmerr, N, Garnero, E. J., 2006. Investigation of upper mantle discontinuity structure beneath the central Pacific using SS precursors. Journal of Geophysical Research 111, B08305, doi:10.1029/2005JB004197.

Schmerr, N., Garnero, E. J., 2007. Upper mantle discontinuity topography from thermal and

chemical heterogeneity. Science 318, 623-626.

Schmid C, Goes S, van der Lee S, Giardini D. 2002. Fate of the Cenozoic Farallon slab from a comparison ofkinematic thermal modeling with tomographic images. Earth Planet. Sci. Lett. 204:17–32

Seno, T., Sakurai, T., Stein, S., 1996. Can the Okhotsk plate be discriminated from the NorthAmerican plate? Journal of Geophysical Research 101, 11305–11315.

Shearer, P. M., 1993. Global mapping of upper mantle reflectors from long-period SS precursors.

Geophysical Journal International 115, 878-904.

Shearer, P. M., 1996. Transition zone velocity gradients and the 520-km discontinuity. Journal of

Geophysical Research 101, 3053-3066.

Shearer, P. M., 2000. Upper mantle seismic discontinuities. Earth’s Deep Interior: Mineral

Physics and Tomography from the Atomic to the Global Scale, Geophysical Monograph 117,

115-131.

Shearer, P. M., Masters, T. G., 1992. Global mapping of topography on the 660 km discontinuity.

Nature 355, 791–796.

1

2

34

5

6789

10

11

121314

15

16171819

20

21

22

23

24

25

26

27

28

29

30

31

32

33

Page 48: University of Alberta › ~ygu › publications › japan_pa… · Web viewTracking Slabs Beneath Northeastern Pacific Subduction Zones Yu Jeffrey Gu University of Alberta, Department

Shearer, P. M., Flanagan, M. F., Hedlin, A. H., 1999. Experiments in migration processing of SS precursor data to image upper mantle discontinuity structure. Journal of Geophysical Research, 104, 7229-7242.

Shen, X., Zhou, H., Kawakatsu, H., 2008. Mapping the upper mantle discontinuities beneath

China with teleseismic receiver functions. Earth Planets Space 60, 713–719.

Shieh, S. R., Mao, H. -K, Hemley, R. J., Chung Ming, L., 1998. Decomposition of phase D in

the lower mantle and the fate of dense hydrous silicates in subducting slabs. Earth and Planetary

Science Letters 159, 13–23.

Smyth, J. R., Frost, D. J., 2001. The effect of water on the 410-km discontinuity: An

experimental study. Geophysical Research Letters 29, doi:10.1029/2001GL014418.

R. Smyth, S. D. Jacobsen, in Earth's Deep Water Cycle, S. D. Jacobsen, S. Van der Lee, Eds. (American Geophysical Union, Washington, DC, 2006), 168, 1–11. 32.

Stixrude, L., Lithgow-Bertelloni, C., Kiefer, B., Fumagalli, P., 2007. Phase stability and shear

softening in CaSiO3 perovskite at high pressure. Phys. Rev. B. 75, 024108.

Suetsugu, D., Inoue, T., Yamada, A., Zhao, D., Obayashi, M., 2006. Towards mappingthe three-dimensional distribution of water in the transition zone fromP-velocity tomography and 660-km discontinuity depths, geophysical monographseries 167, ‘Earth’s deep water cycle’. AGU, 237–249.

Suetsugu, D., Inoue, T., Obayashi, M., Yamada, A., Shiobara, H., Sugioka, H., Ito, A., Kanazawa,

T., Kawakatsu, H., Shito, A., Fukao, Y., 2010. Depths of the 410-km and 660-km discontinuiies

in and around the stagnant slab beneath the Philippine Sea: Is water stored in the stagnant slab?

Physics of the Earth and Planetary Interiors, 183, 270-279.Sugioka, H., Suetsugu, D., Obayashi, M., Fukao, Y., Gao, Y., 2010. Fast P and S wavevelocities associated with the “cold” stagnant slab beneath the northern PhilippineSea. Phys. Earth Planet. Inter. 179, 1–6, doi:10.1016/j.pepi.2010.01.006.

Sung, C. -M., Burns, R. G., 1976. Kinetics of high-pressure phase transformations: Implications

to the evolution of the olivine-spinel transition in the downgoing lithosphere and its

consequences on the dynamics of the mantle. Tectonophysics 31, 1–32.

Tagawa M, Nakakuki T, Tajima F. 2007. Dynamical modeling of trench retreat driven by the slab interaction

1234

5

6

7

8

9

10

11

12

13

141516

171819

20

2122232425

26

27

28

29

30313233

34

35

36

37

3839

Page 49: University of Alberta › ~ygu › publications › japan_pa… · Web viewTracking Slabs Beneath Northeastern Pacific Subduction Zones Yu Jeffrey Gu University of Alberta, Department

with the mantle transition zone. Earth Planets Space 59:65–74

B. Tauzin, E. Debayle, G. Wittlinger (2008), The mantle transition zone as seen by global Pds

phases : no clear evidence for a thin transition zone beneath hotspots, J. Geophys. Res., 113,

B08309, doi:10.1029/2007JB005364.

B. Tauzin, E. Debayle, G. Wittlinger (2010), Seismic evidence for a global low velocity layer in the Earth's upper mantle, Nature Geoscience, 3, 718-721, doi:10.1038/NGEO969.

Tetzlaff, M., Schmeling, H., 2000. The influence of olivine metastability ondeep subduction of oceanic lithosphere. Phys. Earth Planet. Inter. 120,29–38.

Tibi, R., Wiens, D. A., Shiobara, H., Sugioka, H., Yuan, X., 2007. Geophysical Research Letters

34 ,L16316, doi:10.1029/2007GL030527.

Tonegawa, T., Hirahara, K., Shibutani, T., Fujii, N., 2006. Lower slab boundary in the Japan

subduction zone. Earth and Planetary Science Letters 247, 101 – 107.

Torii, Y., Yoshioka, S., 2007. Physical conditions producing slab stagnation: constraintsof the Clapeyron slope, mantle viscosity, trench retreat, and dip angles.Tectonophysics 445, 200–209.

Vacher, P., A. Mocquet, and C. Sotin, Computation of seismic profiles from mineral physics: theimportance of the non-olivine components for explaining the 660 km depth discontinuity, Phys.Earth Planet. Inter., 106, 275-298, 1998.

van der Meijde, M., Marone, F., Giardini, D., van der Lee, S., 2003. Seismic evidence for water

deep in Earth’s upper mantle. Science 300, 1556–1558.

van der Meijde, M., van der Lee, S., Giardini, D., 2005. Seismic discontinuities in the

Mediterranean mantle. Physics of the Earth and Planetary Interiors 148, 233–250.

van der Hilst, R., Engdahl, R., Spakman, W., Nolet, G., 1991. Tomographic imaging of

subducted lithosphere below northwest Pacific island arcs. Nature 353, 37–43.

1

2345

6

789

101112

13

14

15

16

17

18

19202122

232425262728

29

30

31

32

33

34

35

36

37

Page 50: University of Alberta › ~ygu › publications › japan_pa… · Web viewTracking Slabs Beneath Northeastern Pacific Subduction Zones Yu Jeffrey Gu University of Alberta, Department

van der Hilst, R. D., Widiyantoro, S., Engdahl, E. R., 1997. Evidence for deep mantle circulation

from global tomography. Nature 386, 578–584.

van Keken PE, Karato S, Yuen DA. 1996. Rheological control of oceanic crust separation in the transitionzone. Geophys. Res. Lett. 23:1821–24.

Vidale, J. E., Ding, X.-Y., Grand, S. P., 1995. The 410-km-depth discontinuity: A sharp estimate

from near-critical reflections. Geophysical Research Letters 22, 2557-2560.

Wang, Y., Uchida, T., Zhang, J., Rivers, M. L., Sutton, S. R., 2004. Thermal equation of state of

akimotite MgSiO3 and effect of akimotite-garnet transformation on seismic structure near the

660 km discontinuity. Physics of the Earth and Planetary Interiors 143-144, 57-80.

Weidner, D., Wang, Y., 1998. Chemical- and Clapeyron-induced buoyancy at the 660 kmdiscontinuity. J. Geophys. Res. 103, 7431–7441.

Weidner, D., Wang, Y., 2000. Phase transformations; implications for mantle structure.In: Karato, S., Forte, A., Liebermann, R., Masters, G., Stixrude, L (Eds.), Earth's DeepInterior: Mineral Physics and Tomography from the Atomic to Global Scale.American Geophysical Union, pp. 215–235.

Wen, L., Anderson, D. L., 1997. Layered mantle convection: A model for geoid and topography.

Earth and Planetary Science Letters 146, 367 – 377.

Widiyantoro, S., Kennett, B.L.N., van der Hilst, R.D., 1999. Seismic tomography withP and S data reveals lateral variations in the rigidity of deep slabs. Earth Planet.Sci. Lett. 173, 91–100.

B.J. Wood, The effect of H2O on the 410-kilometer seismicdiscontinuity, Science 268 (1995) 74–76.

Yamada, A., Zhao, D., Inoue, T., Suetsugu, D., Obayashi, M., 2009. Seismological evidence for

compositional variations at the base of the mantle transition zone under Japan islands. Gondwana

Research, doi:10.1016/j.gr.2009.04.009

1

2

3

456

7

8

9

10

11

12

13

14

15161718192021222324

25

26

27

28293031

323334

35

36

37

38

Page 51: University of Alberta › ~ygu › publications › japan_pa… · Web viewTracking Slabs Beneath Northeastern Pacific Subduction Zones Yu Jeffrey Gu University of Alberta, Department

Zhao, D., 2004. Global tomographic images of mantle plumes and subductingslabs: insight into deep Earth dynamics. Phys. Earth Planet. Inter. 146, 3–34.Zhao, D., Hasegawa, A., Horiuchi, S., 1992. Tomographic imaging of P and Swave velocity structure beneath northeastern Japan. J. Geophys. Res. 97,19,909–19,928.

Zhao, D., Ohtani, E., 2009. Deep slab subduction and dehydration and their geodynamic

consequences: Evidence from seismology and mineral physics. Gondwana Research 16, 401–

413.

Zhu, G., Shi, Y., Tackley, P., 2010. Subduction of the Western Pacific Plate underneath Northeast

China: Implications of numerical studies. Physics of the Earth and Planetary Interiors 178, 92-99.

Mantle dynamics of Western Pacific and East Asia: Insight fromseismic tomography and mineral physicsDapeng Zhao a, Shigenori Maruyama b, Soichi Omori, Gondwana Research 11 (2007) 120–131.

Zhu, G., Shi, Y., Tackley, P., 2010. Subduction of the Western Pacific Plate underneath Northeast

China: Implications of numerical studies. Physics of the Earth and Planetary Interiors 178, 92 –

99.

Figure Captions:

Fig. 1: (A) A schematic drawing of SS precursor reflection from a subducting oceanic lithosphere

at the base of upper mantle. These waves are sensitive to the depth and impedance contrast of a

mantle interface. (B) Ray theoretical surface reflection points of 6014 high-quality SS waves

used in this study. The main tectonic elements and plate boundaries (Bird, 2003) and slab

contours (Gudmundsson and Smabridge, 1998) are shown by thick and thin black lines,

respectively. The surface projections of five mantle transects (see main text) are labeled A-E,

extending from central Honshu Island (A) to central Kuril Arc (E).

12345

6

7

8

9

10

11

12

13

14

15161718

19

20

21

22

23

24

25

26

27

28

29

30

31

32

Page 52: University of Alberta › ~ygu › publications › japan_pa… · Web viewTracking Slabs Beneath Northeastern Pacific Subduction Zones Yu Jeffrey Gu University of Alberta, Department

Fig. 2: Key steps in the time-to-depth migration of SS precursors. By placing the aligned SS

precursors at the surface (Middle), time samples of transversely polarized seismograms prior to

the arrival of SS (Left) can be effectively mapped to corresponding reflection depths (2nd to the

Right) along the predicted differential time curves based on PREM (Dziewonski and Anderson,

1981). The Right-most panel shows the isotropic shear velocities of PREM down to 1800-km

depth.

Fig. 3: Interpolated reflectivity maps of SS precursor amplitude variations at MTZ (A to C) and

(D) shallow lower mantle depths. An isosurface (threshold = 7.5%) is used to define HRZ in all

panels.. The anomalies marked with green dashed lines are discussed in the text. Slab depth

contours (Gudmundsson and Sambridge, 1998) are drawn by magenta lines at 50 km intervals

from the trench.

Fig. 4: Interpolated CMP gathers along profiles A to D (see Fig. 1A) superimposed on high-

resolution P-wave velocities (Obayashi et al., 2006). Also indicated are earthquakes within the

averaging window of each cross-section. Our interpretations (white lines) are combined with the

-0.5% velocity perturbation contours (red lines, Obayashi et al., 2006).

Fig. 5: (A) Mantle reflectivity structures along the northernmost profile E (see Fig. 1B), P-wave

speeds (Obayashi et al., 2006), and Wadati-Benioff zone seismicity (yellow circles). The thin

red-lines outlines -0.5% velocity perturbations (Obayashi et al., 2006). Our interpretations are

highlighted by the dashed white lines. (B) A schematic interpretation of the HRZs for the Japan-

1

2

3

4

5

6

7

8

9

10

11

12

13

14

15

16

17

18

19

20

21

22

23

Page 53: University of Alberta › ~ygu › publications › japan_pa… · Web viewTracking Slabs Beneath Northeastern Pacific Subduction Zones Yu Jeffrey Gu University of Alberta, Department

Kuril subduction system. The thick red line along the surface of the subducting slab indicates the

ongoing process of dehydration melting. Slab penetration is likely in regions where the 660

appears to be segmented.

1

2

3

4

5

6