nature template - pc word 97€¦  · web viewin all steps of the inversion where phase ... water...

22
1 Supplementary Information: 1.1) Imaging the base of the lithosphere and other discontinuities. The number of events included in the final image (Fig. 2) for each station are as follows: LMN (29), HRV (123), PAL (33), LBNH (50), BINY - north (14), BINY - south (20), SSPA (83). At stations LMN and PAL, only data from back-azimuths of 195-300 and 170-200, respectively, are included to avoid highly complex waveforms on other paths. We present data from BINY in two separate back-azimuthal bins because of the variation in the depth to which the phase from the base of the lithosphere migrates between the events from the north, 110 km, and those from the south, 105 km. The waveforms were migrated using station specific crustal velocity models obtained by modelling scattered waves (Inversion methods are described in more detail in Supplementary Information, section 2 and Supplementary Tables 1 & 2.). For all stations, mantle velocities in the lithosphere were set to representative regional values (see LMN values in Supplementary Table 2) while two lithosphere layers were included for HRV (Supplementary Tables 1 & 2). To evaluate uncertainty in discontinuity depths from lateral variation in mantle velocity, discontinuity depths were also calculated assuming mantle shear-wave velocities beneath each station obtained by surface-wave inversion 1 . At each station, the depth of the lithospheric

Upload: vungoc

Post on 29-Jul-2018

215 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Nature template - PC Word 97€¦  · Web viewIn all steps of the inversion where phase ... water may be assumed to reduce velocity primarily by enhancing seismic attenuation13 using

1

Supplementary Information:

1.1) Imaging the base of the lithosphere and other discontinuities.

The number of events included in the final image (Fig. 2) for each station are as

follows: LMN (29), HRV (123), PAL (33), LBNH (50), BINY - north (14), BINY -

south (20), SSPA (83). At stations LMN and PAL, only data from back-azimuths of

195-300 and 170-200, respectively, are included to avoid highly complex waveforms

on other paths. We present data from BINY in two separate back-azimuthal bins

because of the variation in the depth to which the phase from the base of the lithosphere

migrates between the events from the north, 110 km, and those from the south, 105 km.

The waveforms were migrated using station specific crustal velocity models obtained by

modelling scattered waves (Inversion methods are described in more detail in

Supplementary Information, section 2 and Supplementary Tables 1 & 2.). For all

stations, mantle velocities in the lithosphere were set to representative regional values

(see LMN values in Supplementary Table 2) while two lithosphere layers were included

for HRV (Supplementary Tables 1 & 2).

To evaluate uncertainty in discontinuity depths from lateral variation in mantle

velocity, discontinuity depths were also calculated assuming mantle shear-wave

velocities beneath each station obtained by surface-wave inversion1. At each station, the

depth of the lithospheric discontinuity in general changed by less than 1 km and in no

case changed by more than 1.6 km between the two mantle velocity models. Comparing

this analysis to the very small error bars on discontinuity depth obtained in the formal

inversions (section 2), we infer that the uncertainty in the absolute discontinuity depths

is less than 1.6 km.

When imaging discontinuities in the depth range of the base of the lithosphere,

one must be careful not to mistake reverberated crustal phases for direct converted

Page 2: Nature template - PC Word 97€¦  · Web viewIn all steps of the inversion where phase ... water may be assumed to reduce velocity primarily by enhancing seismic attenuation13 using

2

phases originating from a deeper discontinuity. We verify that the apparent phase from

the base of the lithosphere is a true direct arrival using synthetic seismograms and by

noting that variations in the arrival time of this phase between stations are not correlated

with crustal thickness. This issue is also addressed by migrating waveforms in separate

bins for different ranges of source to station distance, as is shown for station HRV

(Supplementary Fig. 1). Phases that migrate to constant depth with respect to epicentral

distance represent direct conversions, whereas reverberations manifest a false increase

in apparent depth with greater distance. At station HRV, direct converted phases image

the base of the crust (the Moho) at 30 km, a mid-lithospheric phase at 61 km, and a

discontinuity at 97 km which we interpret as the base of the lithosphere due to its

correlation with the approximate thickness of the fast velocity lid seen by surface

waves2,3. Later arrivals are well-modelled as reverberated Moho phases. The depth

predicted for the direct conversion from 97 km varies within the first three epicentral

bins (35 to 50) due to the interference with the first crustal reverberation. Yet,

interference effects in the 45-50 bin are minimal, and including these events in our data

set had no effect on the deconvolved, migrated waveform. Therefore, we omitted data at

distances less than 45 in inversions for the velocity gradient at 97 km.

Although the character of the phases varies between stations, some looking wider

(longer period) than others, the variations are consistent with variations of the entire

impulse response between stations. Therefore, variations between stations in the

character of the phases are more likely due to variable numbers of quality data included

in the bin, the noise level at a particular station, or local structure beneath a station

rather than physical differences in the nature of the discontinuity. A phase from depths

comparable to the base of the lithosphere does appear at station PAL (Fig. 2), although

the migrated waveform signal is noisier than at other stations. Stations SSPA and

LBNH manifest strong phases that are consistent with direct arrivals from depths of 102

km and 110 km, respectively (Fig. 2). Synthetic seismogram modelling shows that the

Page 3: Nature template - PC Word 97€¦  · Web viewIn all steps of the inversion where phase ... water may be assumed to reduce velocity primarily by enhancing seismic attenuation13 using

3

timing and amplitude of these phases may be slightly altered by possible interference

with smaller reverberations from discontinuities internal to the crust. However, although

the possibility of such interference leads to uncertainties in the depth and gradient

associated with a discontinuity at the base of the lithosphere, the discontinuity itself is

still required.

Another receiver function study1 which was located partially in our study region

included analysis of data from station HRV, but the study did not note the existence of a

negative phase at depths near our observed lithosphere-asthensosphere boundary. This

is to be expected given the differences between that study and our own. At HRV, which

is included in both studies, the former study included frequencies up to 1 Hz, whereas

our filters return higher frequencies, up to 2 Hz. Since the high amplitude reverberations

from the Moho arrive just after the phase from the base of the lithosphere

(Supplementary Fig. 1), our testing indicates that filtering higher frequencies from our

data would cause the phase from the lithosphere-asthenosphere boundary to interfere

with the first reverberation from the Moho, losing its appearance as a distinct and

separate phase.

To evaluate the presence of anisotropy in mantle and crustal velocities, we also

considered back-azimuthal groupings of deconvolved and migrated SV and SH

waveforms to search for azimuthal dependence in phase arrival times and amplitudes.

However, the data did not resolve any clear anisotropic signatures.

1.2) Existence of 61 km discontinuity ambiguous.

The 61 km discontinuity observed at station HRV migrates to constant depth with

respect to epicentral distance (Supplementary Fig. 1) which usually indicates that a

discontinuity exists at that depth, instead of the possibility that the phase is a

reverberation from a mid-crustal discontinuity. Although a mid-crustal discontinuity

Page 4: Nature template - PC Word 97€¦  · Web viewIn all steps of the inversion where phase ... water may be assumed to reduce velocity primarily by enhancing seismic attenuation13 using

4

which could be responsible for a reverberation at 61 km is not obvious at HRV where

the crust seems relatively simple, the 61 km phase is not observed consistently at other

stations throughout the region, leading us to question its existence. If a strong mid-

crustal discontinuity does exist beneath HRV, for example, an increase from 5.9 km/s to

6.7 km/s at 11.8 km depth, a negative reverberation could be produced, matching the

magnitude and the depth of the observed 61 km phase. Although such a reverberation

should migrate to increasing depth with respect to epicentral distance, the predicted

move-out is not greater than a difference of 4 km depth from 35 to 80 epicentral degrees

due to the shallow nature of the discontinuity. Therefore, move-out may be difficult to

detect for this case. We have modelled the velocity contrast at the base of the

lithosphere-asthenosphere boundary at HRV without the 61 km discontinuity to

determine the effects of our assumptions regarding the existence of the 61 km

discontinuity, and we find that the contrast at the lithosphere-asthenosphere boundary is

reduced by the absence of a discontinuity at 61 km, but not significantly, to a 3.1-4.9%

velocity contrast (Supplementary Table 2).

2) Inverting for the properties of the base of the lithosphere

Damped least-squares inversions were performed using the migrated waveforms from

our best stations, HRV and LMN, to determine the gradient of the velocity contrast at

the base of the lithosphere. However, we first modelled the structure of each layer and

discontinuity above the lithosphere-asthenosphere boundary, starting with the crust, to

ensure that their effects were appropriately accounted for when inverting for the

properties of the discontinuity at the base of the lithosphere (Supplementary Table 1).

Different observables (e.g. phase timing, amplitude, and waveform shape) were used to

constrain different discontinuities and layer velocities within the model, in order to

maximise the resolution provided by the data while minimising the number of

assumptions regarding structure. Throughout the inversion process, the number of

Page 5: Nature template - PC Word 97€¦  · Web viewIn all steps of the inversion where phase ... water may be assumed to reduce velocity primarily by enhancing seismic attenuation13 using

5

layers in depth was fixed based on observed phases to minimise the number of

assumptions regarding the internal structure of layers. The synthetic seismograms used

to determine partial derivatives were calculated using a propagator matrix method4.

The timing of the Moho Ps phase and the timing of its two, first-order

reverberations in each epicentral bin were used to invert for crustal thickness and

average crustal Vp and Vs (Step 1, Supplementary Table 1). These independent

observations provide excellent resolution of crustal thickness and average velocity.

Such single-layer models over-simplify real crustal structure, and the magnitude of the

velocity increase at the Moho is almost certainly smaller than the values implied by

these crustal velocities. Therefore, we did not include the amplitude of the Moho

conversions in these inversions. Not surprisingly, the amplitudes of the crustal phases

are the one aspect of the synthetic waveforms (red lines in Fig. 2) that do not match the

data (blue lines in Fig. 2). However, we also evaluated the impact that assuming a single

layer crust has on the inferred properties of the lithosphere-asthenosphere boundary and

found that even with very gradational crustal velocities, the velocity drop at the base of

the lithosphere varies by less than 0.5%. For the crust we calculated errors using the

formal 95% confidence limits from the inversion.

Discontinuities below the crust were modelled using waveforms migrated in a

single bin for each station. For HRV, to avoid interference between the phase from the

base of the lithosphere and crustal reverberations for small epicentral distances, we

employed only waveforms from events at distances of 45-80 (Supplementary Fig. 1).

Fortunately, the crustal reverberations at LMN arrive later than at HRV, completely

avoiding interference with the phase from the base of the lithosphere and allowing us to

model data from the full range of epicentral distances represented in the data at LMN

(35-60).

Page 6: Nature template - PC Word 97€¦  · Web viewIn all steps of the inversion where phase ... water may be assumed to reduce velocity primarily by enhancing seismic attenuation13 using

6

In all steps of the inversion where phase amplitudes were modelled (Steps 2 & 3,

Supplementary Table 1), care was taken to accurately calculate predicted migrated

waveform amplitudes and the partial derivatives of the predicted migrated waveform

with respect to discontinuity parameters. In particular, before the real waveforms were

deconvolved, each was normalised by its P-wave component amplitude, and then

weighted according to its signal-to-noise ratio. When the real waveforms were

simultaneously deconvolved and migrated, a water-level, i.e. regularisation, was applied

that in practice removes some of the higher frequency content from signal. To ensure

that the effects of these processing steps were accounted for in the predicted waveform

migration, synthetics were calculated for the P-slowness of each of the real waveforms,

the synthetics were scaled to the amplitudes of their normalised, real waveform

counterparts, and identical water-levels were applied. This procedure also ensured that

we accurately accounted for differences in phase amplitude produced by waveforms

with varying P-wave incidence angle.

When matching deconvolved, migrated waveforms in depth (Steps 2 & 3,

Supplementary Table 1), it is important that the migration model not bias the results.

Hence, we required the migration model to match our best-fitting model. To accomplish

this, the parameters that we held fixed in the inversion were held fixed at identical

values in the migration model. The parameters that varied during the inversion were

updated in the migration model once several iterations were completed and convergence

was achieved. Then the data were reprocessed using the new migration model, and the

iterative inversion was re-run. The process was repeated until subsequent inversions no

longer yielded perturbations to the best-fitting model. The number of layers in the

migration model remained constant throughout all steps of the inversion. Predicted

waveforms and partial derivatives were re-calculated for each iteration.

Page 7: Nature template - PC Word 97€¦  · Web viewIn all steps of the inversion where phase ... water may be assumed to reduce velocity primarily by enhancing seismic attenuation13 using

7

We inverted the timing and amplitude of the mid-lithospheric phase observed at

HRV to determine the depth of the potential associated discontinuity and the magnitude

of the velocity contrast (Step 2, Supplementary Table 1). Vp/Vs in the mantle

lithosphere was fixed at 1.8, and Vp in the layer above the mid-lithospheric

discontinuity was held at 8.2 km/s as determined by local refraction experiments for

depths just below the Moho5. A 5.4 ± 0.6 % shear velocity decrease with depth at 60.9 ±

0.4 km depth best fits the observed mid-lithospheric phase at HRV (Supplementary

Table 2). Errors correspond to the change in model parameter for which the migrated

synthetic waveform image hits the bounds of the two standard deviation error bars for

the depth and amplitude of the observed phase; the latter were calculated using a

bootstrap method similar to that described in Fig. Legend 2. At LMN, mantle velocity

was assumed to be constant from the Moho to the base of the lithosphere, again

matching the sub-crustal lithosphere velocity determined by refraction data5. We also

modelled HRV without the mid-lithospheric discontinuity at 61 km depth due to some

ambiguity regarding its existence (Supplementary Information, section 1.2 and

Supplemental Table 2).

At the lithosphere-asthenosphere boundary we used the complete waveform shape

to constrain not only the depth and total velocity contrast but also the thickness of the

velocity gradient assuming a linear gradient (Step 3, Supplementary Table 1).

Waveforms converted at the lithosphere-asthenosphere boundary depend on four

parameters: the velocity contrast at the discontinuity, the depth range over which the

velocity contrast occurs, the absolute depth of the discontinuity, and the dominant

period of the incident P-wave. A variety of tests including forward modelling, grid

searches, and formal inversions indicate strong trade-offs between the parameters.

Fortunately, the dominant period of the incident P-wave may be constrained

independently. Limiting the acceptable range of incident P-wave period in turn limits

the possible variation in the other parameters.

Page 8: Nature template - PC Word 97€¦  · Web viewIn all steps of the inversion where phase ... water may be assumed to reduce velocity primarily by enhancing seismic attenuation13 using

8

To determine the dominant period of the incident P-wave we simultaneously

deconvolved the P-wave signals from themselves, applying the same normalization and

water-levels that are applied to the Ps data. The result is an impulse response with a

width that represents the dominant period of the incident P-waves in the data. We then

simultaneously deconvolved synthetic P-waves that matched the p-slownesses and

amplitudes of the data and applied the same water-levels, assuming a single period for

all of the simultaneously deconvolved synthetic P-waves in a given impulse response.

We created synthetic auto-deconvolved impulse responses for a variety of incident P-

wave periods until we found the auto-deconvolved impulse response that fit the data

best. The 95% confidence limits for the best-fitting incident P-wave period at each

station were established using an F-test. The dominant periods that fall within the 95%

confidence limits for HRV are 1.5-2.0 s, and those for LMN are 2.9-3.7 s.

Because parameter trade-offs are strong, and to avoid the exclusion of viable

models, we fixed gradient thickness at a variety of values and inverted for the best-

fitting period, depth, and velocity contrast. Supplementary Fig. 2 shows the best-fitting

velocity drops and P-wave periods that correspond to the fixed gradient thicknesses.

The misfits between the data and the deconvolved, migrated Ps waveforms for these

parameter combinations are approximately equal, increasing slightly for the 12 km thick

gradient at LMN. However, we accept only those models for which the incident P-wave

periods fall within the 95% confidence limit for the best-fitting incident P-wave period

(Supplementary Fig. 2b). At HRV a 3.3-5.7% shear velocity drop over 5 km or less at

96.6-96.8 km depth is acceptable at 95% confidence, assuming the existence of the mid-

lithospheric discontinuity. Without the mid-lithospheric discontinuity, the velocity drop

is 3.1-4.9%. At LMN a 6.8-7.4% velocity contrast that occurs over 5 km or less at 90.6-

90.7 km depth gives the best fit to the dominant period of the incident waveform. Yet, a

velocity contrast that occurs over 10 km requiring a 10.7% contrast cannot be excluded

by the 95% confidence limit on the dominant period of the incident P-wave at LMN.

Page 9: Nature template - PC Word 97€¦  · Web viewIn all steps of the inversion where phase ... water may be assumed to reduce velocity primarily by enhancing seismic attenuation13 using

9

The error bars on the absolute depth of the discontinuity reported here are the formal

error bars from the inversion. However, we prefer to use the broader 1.6 km error bars

inferred by comparing the absolute depths obtained with this approach to those obtained

assuming surface wave velocity structure beneath each station2 (Supplementary

Information, section 1.1).

During the inversions for the parameters above, density and Vp/Vs were assumed

to be constant across the lithosphere-asthenosphere boundary, at 3.32 g/cm3 and 1.8

respectively. However, we also performed inversions in which the density decreased by

10% with depth and Vp/Vs ratio rose from 1.80 to 1.85 with depth over the gradient of

the discontinuity. Reducing the density at the base of the lithosphere by 10% has only a

small effect on the velocity contrast, decreasing it by ~0.25%. Increasing Vp/Vs by 0.05

affects the P-wave velocity contrast dramatically, reducing it by ~2.75% relative to the

value determined for fixed Vp/Vs; however, the necessary S-wave velocity contrast

remains unaffected. Overall, our assumptions concerning density and Vp/Vs do not

significantly affect our conclusions regarding S-wave velocity properties at the

discontinuity.

We assumed a linear velocity gradient at the lithosphere-asthenosphere boundary.

However, we did experiment with exponential gradients, and for gradients over the

same depth range, the magnitude of the velocity contrasts required are nearly identical

for the two cases. Hence, we cannot significantly modify our results by changing the

shape of the velocity gradient.

3) Expected velocity gradients due to variations in temperature, hydration, grain

size, and melt.

Page 10: Nature template - PC Word 97€¦  · Web viewIn all steps of the inversion where phase ... water may be assumed to reduce velocity primarily by enhancing seismic attenuation13 using

10

If the observed velocity contrast is caused by purely thermal gradients, olivine

deformation experiments6 suggest that it can be described according to the following

equations:

∂lnVS/∂T = ∂lnVSU/∂T - F(α) [Q-1(ω,T)/π] (E+PV)/RT2

where

Q-1=Ad-mToαexp(-α(E+PV)/RT)

F(α)=(πα/2)cot(πα/2)

and the change in the unrelaxed shear wave speed with respect to temperature,

∂lnVSU/∂T = -8.6x10-5 K-1; activation energy, E=424,000 J/mol; A=730 s-αµmm; grain

size, d=1mm; the frequency (ω) dependence of attenuation, α=.26; the grain size

sensitivity of attenuation, m=.26; and activation volume, V=6x10-6 m3/mol (refs 6, 7) .

Since some of these parameters were determined by matching data at experimental

conditions, we subtract the 200 MPa pressure at which the experiments were performed6

from the pressure at 100 km depth, 3.36 GPa, and use the result as pressure, P. For the

period, To, we use the dominant incident P-wave periods which fall within our

confidence limits, 1.5 and 2 s at HRV, corresponding to 5.7 and 3.3% velocity contrasts

respectively, and 2.9 and 3.7 s at LMN, corresponding to 10.7 and 6.8% velocity

contrasts respectively. Since an increased absolute background temperature, T,

decreases the thermal contrast required to explain a given velocity contrast, we have

chosen a relatively high asthenospheric temperature to be conservative. We find that at

1375° C we require at least a 119-210° C thermal contrast to explain the 3.3-5.7%

velocity contrast at HRV and a 222-374° C thermal contrast to explain the 6.8-10.7%

velocity contrast at LMN. Other parameter choices are also conservative. For example,

the value of 6x10-6 m3/mol for activation volume is at the low end of experimentally

determined values8, and larger activation volumes would increase the required

Page 11: Nature template - PC Word 97€¦  · Web viewIn all steps of the inversion where phase ... water may be assumed to reduce velocity primarily by enhancing seismic attenuation13 using

11

temperature contrast. Similarly, 1 mm is small for olivine grain size at temperatures and

stresses appropriate for ~100 km depth9, and a greater mantle grain size (say 10 mm)

would also require a larger temperature jump (173-315° C at HRV and 331-552° C at

LMN). Alternative physical models10 have been proposed to explain the experimental

data6 we employ; however, these alternative interpretations would not significantly

affect the thermal gradients we calculate.

In numerical models of mantle flow and temperature that allow for small-scale

convection at the transition from thick cratonic lithosphere to thinner lithosphere in

which viscosity is dependent solely on temperature and pressure11,12, the thermal

gradients at the base of the lithosphere are typically smaller than 2-5° C/km. Even

allowing for untested combinations of model parameters, 10° C/km is a generous upper

bound on the thermal gradient possible at the base of a ~100 km thick lithosphere12. A

gradient of 5° C/km would predict at most 55° C over 11 km at LMN and 25° C over 5

km at HRV. Using the parameter values described in the previous paragraph, a 55° C

thermal contrast results in a 1.7% maximum velocity contrast for the dominant periods

at stations LMN, and a 25° C thermal contrast results in a 0.7% maximum velocity

contrast for the dominant periods at HRV. Such velocity contrasts are far too small to

explain the 3.1-10.7% velocity drop required by our data. A gradient of 10° C/km would

produce a 3.3% maximum velocity contrast for the dominant periods at stations LMN,

and a 1.3% velocity contrast for the dominant periods at HRV. These predictions still

fail to match the minimum velocity drop required at LMN (6.8%) and HRV (3.1%).

Additionally, these mantle flow models allow significant small-scale convection

to develop only in situations where the upper plate is relatively stationary with respect

to the asthenosphere12. If shearing due to the motion of the North American continental

lithosphere over the asthenosphere is allowed, the thermal gradient at the base of the

Page 12: Nature template - PC Word 97€¦  · Web viewIn all steps of the inversion where phase ... water may be assumed to reduce velocity primarily by enhancing seismic attenuation13 using

12

lithosphere will be reduced12, thus reinforcing the statement that thermal contrasts alone

are insufficient to explain the velocity contrast observed in our data.

The magnitude of a shear-wave velocity drop due to a dehydration front at the

lithosphere-asthenosphere boundary is difficult to estimate due to a lack of direct

experimental measurements of the effects of water on olivine and other upper mantle

minerals. However, water may be assumed to reduce velocity primarily by enhancing

seismic attenuation13 using the following relationship14:

V (ω, T, P, C) =V0 (T, P) [1-1/2cot (πα/2) Q-1(ω, T, P, C)]

where V0(T,P) is the anharmonic seismic wave velocity, ω is frequency, T is

temperature, P is pressure, and C is composition. Given that the shear-wave velocity

drop at the lithosphere-asthenosphere boundary and the asthenospheric velocities we

observe are comparable to the values for old Pacific Ocean regions15, we used values of

shear-wave Q (inverse attenuation) for the lithosphere (150) and the asthenosphere (50)

from this study15 to estimate the effects of weakly frequency dependent attenuation on

velocity. Assuming values of α, a parameter that describes the frequency dependence of

attenuation, of 0.1-0.3, water would reduce shear-wave velocity by 1.3- 4.3%. Thus, the

velocity drop that can be attributed to the presence of water overlaps the minimum

velocity contrast at HRV (3.1%) but does not attain the minimum value at LMN (6.8%).

What is the role of grain size? Grain size, like water, affects the attenuation

component of velocity6. Therefore, the total effect of water and grain size on velocity is

also limited by observed values of attenuation to be < 4.3%. Could grain size variation

contribute to sharpening the velocity contrasts predicted from purely thermal models?

This scenario is very unlikely. Since grain size is inversely proportional to stress16-18, a

sharp increase in shear stress with depth would be required for grain size to be

responsible for the observed velocity contrast. However, in models in which viscosity

Page 13: Nature template - PC Word 97€¦  · Web viewIn all steps of the inversion where phase ... water may be assumed to reduce velocity primarily by enhancing seismic attenuation13 using

13

depends only on temperature and pressure12, shear stress generally decreases gradually

with depth.

Alternatively, if asthenospheric temperatures are above the damp-solidus, melt

fractions of 1.4-1.8% (ref. 19) or 3.6-5.7% (ref. 20) in the asthenosphere could easily

produce the roughly 7% velocity drop inferred for the best-fitting LMN models and

even the 11% velocity drop associated with the largest gradient thickness permitted by

the LMN data. In scenarios involving melt, the base of the lithosphere would be defined

by the damp-solidus, thus prohibiting melt from rising into the lithosphere. For

peridotite solidi corresponding to 800-1000 H/106 Si (ref. 21), a range of reasonable

mantle temperatures will be above the solidus at asthenospheric depths, and will cross to

temperatures below the solidus at the 90-110 km depths we infer for the lithosphere-

asthenosphere boundary. One possible model for generating melt in the asthenosphere

beneath eastern North America would be upward flow and decompression of mildly

hydrated asthenospheric material along the contours of the more rigid, shallowing

lithosphere. This type of flow occurs in mantle flow models that incorporate the WSW

absolute motion of the North American plate12,22 (Fig. 1). Alternatively, mantle

temperatures may exceed the solidus within the asthenosphere without the need for

decompression21. In general, if melt leaves the region, drawing the water that has

partitioned into it away, a mechanism to rehydrate stationary mantle or bring in new

hydrated mantle is required. Such re-supply of hydrated mantle is provided by the

upward flow beneath the continental interior described above. In any model involving

wet melting, and in the earlier hypothesis that involves a sub-solidus hydrated

asthenosphere, a reasonable question is how the diffusion of hydrogen into the

postulated dry sub-solidus lithosphere would affect lithospheric properties. The

presence of hydrogen would effectively hydrate and locally reduce velocities, but the

maximum length scale for this diffusion over the 300-450 My since emplacement of the

Page 14: Nature template - PC Word 97€¦  · Web viewIn all steps of the inversion where phase ... water may be assumed to reduce velocity primarily by enhancing seismic attenuation13 using

14

lithosphere is 6-7 km (ref. 21), consistent with the sharp lithosphere-asthenosphere

boundary imaged in this study.

1. Li, A., Fischer, K. M., van der Lee, S. & Wysession, M. E. Crust and upper mantle discontinuity structure in eastern North America. J. Geophys. Res. 107, doi: 10.1029/2001JB000190 (2002).

2. Li, A., Forsyth, D. W. & Fischer, K. M. Shear velocity structure and azimuthal anisotropy beneath eastern North America from Rayleigh wave inversion. J. Geophys. Res. 108, doi: 10.1029/2002JB02259 (2003).

3. van der Lee, S. High-resolution estimates of lithospheric thickness from Missouri to Massachusetts, USA. Earth Planet. Sci. Lett. 203, 15-23 (2002).

4. Keith, C. M. & Crampin, S. Seismic body waves in anisotropic media; synthetic seismograms. Geophys. J. Royal Astron. Soc. 49, 225-243 (1977).

5. Musacchio, G., Mooney, W. D., Luetgert, J. H. & Christensen, N. I. Composition of the crust in the Grenville and Appalachian provinces of North America inferred from Vp / Vs ratios. J. Geophys. Res. 102, 15,225-15,241 (1997).

6. Jackson, I., Fitz Gerald, J. D., Faul, U. H. & Tan, B. H. Grain size sensitive seismic wave attenuation in polycrystalline olivine. J. Geophys. Res. 107, doi: 10.1029/2001JB001225 (2002).

7. Issak, D. G. High-temperature elasticity of iron-bearing olivines. J. Geophys. Res. 97, 1871-1885 (1992).

8. Hirth, G. & Kohlstedt, D. L. in Inside the Subduction Factory (ed. Eiler, J.) 83-105 (Geophysical Monograph 138, American Geophysical Union, Washington DC, 2003).

9. Evans, B., Renner, J. & Hirth, G. A few remarks on the kinetics of static grain growth in rocks. Int. J. Earth Sci. 90, 88-103 (2001).

10. Cooper, R. F. in Plastic Deformation of Minerals and Rocks (ed. Wenk, H. R.) 253-290 (Mineralogical Soc. America, Rev. Mineal. Geochem. 51, Washington DC, 2002).

11. King, S. D. & Ritsema, J. African hot spot volcanism: small scale convection in the upper mantle beneath cratons. Science 290, 1137-1140 (2000).

12. Zaranek, S. E., Parmentier, E. M. & Fischer, K. M. Effects of basal drag and ablation on the evolution of cratonic lithosphere. Eos Trans. AGU, T41C-03 (2004).

13. Karato, S.-I. & Jung, H. Water, partial melting and the origin of the seismic low velocity and high attenuation zone in the upper mantle. Earth Planet. Sci. Lett. 157, 193-207 (1998).

14. Minster, J. B. & Anderson, D. L. A model of dislocation-controlled rheology for the mantle. Philos. Trans. R. Soc. London A 299, 319-356 (1981).

15. Gaherty, J. B., Kato, M. & Jordan, T. H. Seismological structure of the upper mantle; a regional comparison of seismic layering. Phys. Earth Planet. Inter. 110, 21-41 (1999).

16. Twiss, R. J. Theory and applicability of a recrystallized grain size paleopiezometer. Pure and Applied Geophysics 115, 227-244 (1977).

Page 15: Nature template - PC Word 97€¦  · Web viewIn all steps of the inversion where phase ... water may be assumed to reduce velocity primarily by enhancing seismic attenuation13 using

15

17. Karato, S. I., Toriumi, M. & Fujii, T. Dynamic recrystallization of olivine single crystals during high-temperature creep. Geophys. Res. Lett. 7, 649-652 (1980).

18. Van der Wal, D., Chopra, P., Drury, M. & Fitz Gerald, J. Relationships between dynamically recrystallised grain size and deformation conditions in experimentally deformed olivine rocks. Geophys. Res. Lett. 20, 1479-1482 (1993).

19. Hammond, W. C. & Humphreys, E. D. Upper mantle seismic wave velocity; effects of realistic partial melt geometries. J. Geophys. Res. 105, 10975-10986 (2000).

20. Kreutzmann, A. et al. Temperature and melting of a ridge-centred plume with application to Iceland. Part II: Predictions for electromagnetic and seismic observables. Geophys. J. Int. 159 (2004).

21. Hirth, G. & Kohlstedt, D. L. Water in the oceanic upper mantle; implications for rheology, melt extraction and the evolution of the lithosphere. Earth Planet. Sci. Lett. 144, 93-108 (1996).

22. Fouch, M. J., Fischer, K. M., Parmentier, E. M., Wysession, M. E. & Clarke, T. J. Shear wave splitting, continental keels, and patterns of mantle flow. J. Geophys. Res. 105, 6255-6275 (2000).

Supplementary Fig. 1

SV waveforms from HRV deconvolved and migrated in epicentral distance bins.

Red signifies positive polarity, or a velocity increase with depth, while blue

indicates a negative amplitude corresponding to a velocity decrease with depth.

Black lines indicate the migrated depths of synthetic phases predicted from our

preferred HRV model, including direct conversions from the Moho (30 km), 61

km, and 97 km depth. Later arrivals show greater apparent depth with distance

and correspond to crustal reverberations.

Supplementary Fig. 2

Parameter trade-offs and model constraints. Blue lines represent LMN models,

while green lines show HRV results. Red stars indicate models that fit the Ps

conversion from the base of the lithosphere, and whose dominant incident P-

wave periods fall within the 95% confidence limits from the auto-deconvolution

test. Black stars are models that fit the Ps conversion from the base of the

Page 16: Nature template - PC Word 97€¦  · Web viewIn all steps of the inversion where phase ... water may be assumed to reduce velocity primarily by enhancing seismic attenuation13 using

16

lithosphere, but require P-wave periods outside the 95% confidence limits. a)

The best-fitting models for several gradient thicknesses show that the observed

Ps conversion from the base of the lithosphere can be fit with a variety of

models at HRV and LMN. However, each model corresponds to a single

dominant incident P-wave period, and the dominant period may be constrained

independently. b) RMS misfit of the auto-deconvolved P-wave synthetics to the

auto-deconvolved P-wave data for the dominant periods used with the models

in Supplementary Fig. 2-a. The numbers next to the stars indicate the gradient

thickness that corresponds to that P-wave period. The red dashed lines show

the 95% confidence limits on the dominant period at HRV and LMN, indicating

that models with gradients as thick as roughly 5 km are acceptable at HRV,

while at LMN the gradient may occur over 11 km or less.