04.08.2013 Views

A marine wide angle reflection/refraction profile adjacent Nova ...

A marine wide angle reflection/refraction profile adjacent Nova ...

A marine wide angle reflection/refraction profile adjacent Nova ...

SHOW MORE
SHOW LESS

Create successful ePaper yourself

Turn your PDF publications into a flip-book with our unique Google optimized e-Paper software.

Crustal Structure of the Meguma Terrane<br />

Offshore of <strong>Nova</strong> Scotia and its Tectonic<br />

Implications<br />

Version Dec 11/2006<br />

by<br />

H. R. Jackson 1 , D. Chian 1 , K. Louden 2 and M. Salisbury 1<br />

1 Geological Survey of Canada (Atlantic), Box 1006, Dartmouth, NS, Canada B2Y 4A2<br />

2 Department of Oceanography, Dalhousie University, Halifax, <strong>Nova</strong> Scotia, Canada B3H 4J1<br />

ESS contribution #<br />

E-mail: rujackso@nrcan.gc.ca; phone: 902-426-3791; fax: 902-426-6152


Abstract<br />

A 500 km long <strong>marine</strong> <strong>wide</strong>-<strong>angle</strong> <strong>reflection</strong>/ <strong>refraction</strong> (WAR) line was run parallel to the coast<br />

of <strong>Nova</strong> Scotia, on the Meguma Terrane. The terrane’s origin has been the subject of controversy<br />

exacerbated by the lack of exposure of lower crustal rocks. To provide constraints on the crust, P- and S-<br />

wave velocity models were developed using the data from 14 ocean bottom seismometers. Limited<br />

vertical incidence <strong>reflection</strong> <strong>profile</strong>s provide control for the modelling process. In situ laboratory<br />

measurements of the velocities and densities of rocks that are possible candidates for the crust were<br />

compared with the WAR velocities. The near seabed layer with velocity of 5.5-5.7 km/s and a thickness of<br />

3 km is interpreted as the Meguma Supergroup. A 10 km thick 4.3-5.0 km/s layer observed in the<br />

Orpheus Graben is inferred to be sedimentary strata. The velocity of the crust is 6.0-6.6 km/s, with a<br />

Poisson’s ratio of = 2.4 indicative of a felsic composition. Reflection <strong>profile</strong>s show a transparent upper<br />

crust that can be correlated with granitoid plutons. The lowermost crust is highly reflective. Both the<br />

Liscomb Complex and the Tangier granulite xenolith suite have high and low velocity members that could<br />

be the cause of the reflectivity. The subsurface contact between the Meguma and Avalon terranes is a<br />

southerly facing ramp with a length of about 60 km. The South Mountain Batholith of the Meguma<br />

Terrane is located to the southeast of the subsurface extent of the Avalon Terrane. Several hypotheses for<br />

the emplacement of the Meguma Terrane are examined. Delamination is one model that is consistent with<br />

the new constraints.<br />

2


Introduction<br />

The Meguma Terrane (Fig. 1) consists principally of metasedimentary rocks of the Meguma<br />

Supergroup intruded by peraluminous granites. The basement to the extensive Cambro-Ordovician<br />

metasedimentary rocks (Fig. 2) is not exposed and its type, origin, and emplacement history have been<br />

difficult to determine (Greenough et al. 1999). The Meguma Terrane was the last to be added to the<br />

Appalachian Orogen, <strong>adjacent</strong> to the Avalon Terrane (Williams and Hatcher 1982). Keppie and<br />

Dallmeyer (1995) and Eberz et al. (1991) have suggested that the basement is Precambrian and of African<br />

origin or that the Meguma Terrane is thrust over the high-grade gneisses of the Avalon Terrane<br />

(Greenough et al. 1999). Murphy et al. (1999) hypothesized that hot spot activity was the cause of the<br />

voluminous granites and lesser mafic rocks. Pe-Piper and Jansa (1999) based on geochemical evidence<br />

advocate that the lower crust of the Meguma and Avalon terranes are similar and the voluminous igneous<br />

activity that produced the granites was produced by post-orogenic regional extensional. The available<br />

<strong>refraction</strong> data (Barrett et al. 1964) indicate a low overall crustal velocity incompatible with the Meguma<br />

Terrane overlying a high velocity basement due to Avalon crust or mafic underplating.<br />

The scientific objectives of this WAR experiment were 1) to develop a velocity model for the<br />

Meguma Terrane and 2) to determine the velocity of and depth relationship to the <strong>adjacent</strong> Avalon<br />

Terrane to gain insights into the enigmatic evolution of the Meguma Terrane. Bedrock samples, swath<br />

bathymetry and potential field data extend the surface geology offshore to aid in the interpretation of our<br />

WAR data. Limited seismic <strong>reflection</strong> <strong>profile</strong>s are available to compare the <strong>reflection</strong> characteristics of<br />

the crust to the WAR data. Significantly, our model velocities are evaluated against the velocities from<br />

3


laboratory measurements of candidate rocks for the crust. Gravity modelling is done to provide additional<br />

control on the velocity model.<br />

Geological and Geophysical Framework<br />

The Meguma Supergroup (Fig. 1) consists mostly of fine-grained siliciclastic turbidites and is of<br />

Gondwanan affinity based on its dispersal pattern (Schenk 1970, 1971). Its composite stratigraphic<br />

thickness exceeds 23 km but nowhere do all the units occur (Schenk 1995; 1997). It was deposited during<br />

the Cambrian (or older) to the Early Ordovician (Fig. 2). The Supergroup has a monotonous lithotectonic<br />

assemblage with similar structural style and level of exposure across the belt. There is no preservation of<br />

shelfal platform carbonate sequences and no exposure of deep-seated rocks. Deformation in the<br />

Supergroup was accommodated by chevron folding.<br />

Intruding it during from 380-370 Ma ( Keppie and Dallymeyer 1995) and representing at least<br />

25% of the exposed rocks are the peraluminous granites (Clark and Chatterjee 1992). The intrusions are<br />

massive; for instance, the South Mountain Batholith (SMB) (Fig. 1) is the largest in the Appalachian<br />

Orogen (Benn et al. 1997). Mafic intrusions are found only in the peripheral plutons and may contribute to<br />

the forming of the tonalite (Clarke et al. 2000) but are not a significant proportion of the outcrop. The<br />

mid-lower crust must have experienced high T/low P metamorphism to produce large volumes of<br />

granitoid rocks. It is unclear how to reconcile the history of the metasedimentary rocks with the extensive<br />

melting of the crust.<br />

Only three windows into the rocks of the deeper crust are available (Clarke at al. 1993): the<br />

granitoid intrusions; the Liscomb Complex consisting of granites, mafics, ortho-and paragneiss; and a<br />

4


small group of lamprophyre dykes at Tangier, containing xenoliths, felsic and mafic granulites (Fig. 1).<br />

The composition of the granitoid intrusions indicates that the Meguma Supergroup cannot be the only<br />

source for the intrusions and there must be changes in rock type at depth (Clarke et al. 1988).<br />

Furthermore, the Sm/Nd systematics on the Devonian granites suggest they were formed from deep<br />

crustal material that is significantly younger than the Cambro-Ordovician metasedimentary rocks they<br />

intrude (Clarke et al. 1988). This age relationship reveals an inverted stratigraphy for the Meguma<br />

Terrane.<br />

The metamorphic grade of the Liscomb Complex is higher (amphibolite to granulite) than the<br />

Meguma Supergroup, but lesser than the lamprophyre dykes (Clarke et al. 1993), suggesting that it<br />

represents an intermediate level in the crust. Felsic xenoliths from the Tangier dyke (Fig. 1) have the<br />

appropriate composition to produce the granites. Their Tchur model ages are younger than the Meguma<br />

Supergroup flysch (Eberz et al. 1991) but consistent with the age of the granites.<br />

The Meguma Terrane was accreted to North America during the Acadian Orogeny in the Early<br />

Devonian (ca 400 Ma; Fig. 2) along the Cobequid-Chedabucto Fault (CCF) (Withjack et al. 1995,<br />

Webster et al. 1998, Pe-Piper and Piper, 2004). This fault can be traced from offshore Cape Breton across<br />

<strong>Nova</strong> Scotia into the Bay of Fundy. The CCF is exposed onshore and mapped using various geological<br />

and potential field data. This combined data set indicates that from surface to 6 km depth the fault is steep<br />

(Webster et al. 1998). Offshore of Cape Breton (Fig. 1) at the end of line 99-1 (Fig. 3), seismic <strong>reflection</strong><br />

<strong>profile</strong> 89-1 (Marillier et al. 1989) shows the Avalon-Meguma contact as southward dipping.<br />

To place spatial limits on the Meguma Terrane the potential field and seismic information are<br />

examined. The regional gravity (Dehler and Roest 1998; Fig. 3) and magnetic anomaly maps (not shown;<br />

5


see Oakey and Dehler 1998) cover both the onshore and offshore. Onshore the potential field information<br />

is readily compared with the surface geology. It is extrapolated offshore by swath bathymetry, samples of<br />

basement from petroleum wells and short hard rock cores. Several regional negative gravity anomalies are<br />

observed in proximity to WAR 99-1 <strong>profile</strong>. The two largest are of similar shape and extent, one is<br />

centred on <strong>Nova</strong> Scotia and continues to the southern part of line 1 and the other is observed entirely<br />

offshore. The first anomaly coincides with the SMB and its near shore extent (Fig. 3) based on swath<br />

bathymetric mapping (Loncarevic et al. 1994). The major offshore gravity low is correlated with granitic<br />

rocks based on samples from D-26 and E-07 wells. A linear negative anomaly south of Cape Breton<br />

Island is associated with the thick accumulation of sedimentary rocks in the Orpheus Graben, the offshore<br />

extension of the CCF.<br />

Onshore the distribution of the Meguma Supergroup correlates with linear magnetic anomalies.<br />

This is due to mineralization along axial folds that occur evenly at about 8 km spacing. In contrast to the<br />

clastic rocks, the South Mountain and Musquodoboit batholiths that have similar mineralogy, petrology<br />

and geochemistry (Ham 1998) produce broad negative anomalies. Offshore the magnetic pattern<br />

associated with the Meguma Supergroup can be followed to the location of line 1 and beyond. This is<br />

substantiated by multibeam soundings that show linear bed forms associated with the Meguma<br />

Supergroup (Loncarevic et al. 1994). In addition, scientific drilling recovered a quartzite petrologically<br />

typical of the Meguma Supergroup (Pe-Piper and Loncarevic 1989) near the southwestern section of line<br />

99-1 (Fig. 3 Core 22). Further offshore the extent of the Supergroup is confirmed by phyllites and<br />

metaquartzites drilled in the offshore wells I-94, N-30, I-22, F-38 and E-53 (Fig. 3) (Jansa 1991; Pe-Piper<br />

and Jansa 1999).<br />

6


Three deep multichannel seismic <strong>reflection</strong> lines sampled on the Meguma Terrane on the margin<br />

(Keen et al. 1991a, b; Keen and Potter 1995). The <strong>profile</strong>s all showed a depth to the <strong>reflection</strong> Moho of<br />

between 10-12 s that is less than the <strong>adjacent</strong> Avalon Terrane at 12-14 s. Profile 86-5B (Marillier et al.<br />

1989) reveals the contact of the thicker Avalon crustal block with the Meguma Terrane south of the<br />

Orpheus Graben. Southward dipping reflectors cut the crust to the Moho separating the two crustal<br />

blocks. The cross-section of the crust on the Scotian shelf is displayed on deep multichannel <strong>reflection</strong><br />

<strong>profile</strong> 88-1 (Keen et al. 1991b) and its extension 89-2 (Fig. 4). A transparent upper crust is observed to a<br />

depth of 15 km that is superseded by a moderately reflective crust to a depth of 25-30 km. The lower crust<br />

is most reflective from 33-38 km depth.<br />

None of the <strong>reflection</strong>s can be correlated with the previous <strong>refraction</strong> results of Barrett et al.<br />

(1964). They measured velocities of 5.5, 6.10 and 8.11 km/s for the metasedimentary layer, crust and<br />

upper mantle. The small number of recording stations 4 and the shot spacing of 15 km suggested that the<br />

improvements in technology and interpretation techniques in the intervening 40 years would produce a<br />

higher resolution velocity depth <strong>profile</strong> if the line was rerun.<br />

Experiment<br />

On Hudson cruise 99-007, 14 digital ocean bottom seismometers (OBS) were launched and<br />

recovered (Fig. 3). The airgun array fired 3983 shots over a range of 503 km. The data quality was<br />

reasonable, except on the noisy OBS 13 deployed near the Orpheus Graben. OBS 1-8 recorded only about<br />

75% of the data due to a computer-programming fault. A second line (99-2) was positioned nearly<br />

perpendicular to the first (Jackson et al. 2000).<br />

7


The source for both lines was an array of 6 airguns with a total capacity of 6x16.4 L (6 x 1000 in 3 )<br />

deployed at 20 m below sea surface and fired every 60 s at 4 knots for a shot interval of ~110 m. Global<br />

Positioning System was used to determine the shot positions. The untuned array produces a reverberatory<br />

signal; however, this signal has the high energy at 2-10 Hz needed for long distance propagation and<br />

develops shear waves.<br />

Shot times, corrected for clock drift in both the OBS and the firing system for the airguns, were<br />

combined with navigation to generate a shot table for each OBS. This in turn was used to convert the raw<br />

OBS data to SEGY format at sea. We process the SEGY data with a causal band-pass filter of 3-10 Hz.<br />

The data are displayed with gain increasing linearly with shot-receiver offset. When data quality are not<br />

the best, we apply coherency mixing to the data similar to that described by Chian and Louden (1992): an<br />

average of ~7 <strong>adjacent</strong> traces along a coherency-determined optimal slope between 2 and 8 km/s, which<br />

results in minimal distortion of wavelet. Due to the large number of OBS <strong>profile</strong>s to be presented, the<br />

WAR data are picked from the hydrophone (and occasionally geophone) channels, the height of which<br />

represent the estimated travel-time uncertainty, overlain with computed travel-time curves. Unfiltered data<br />

were sometimes used during digitization of the water waves and <strong>refraction</strong>s from shallow sediment given<br />

their higher frequency content relative to deeper phases. For weak phases, such as some Moho <strong>reflection</strong>s<br />

and <strong>refraction</strong>s from lower crust and upper mantle, coherency mixing is found useful to properly identify<br />

arrival times and digitize them.<br />

Modelling<br />

The two-dimensional ray-tracing algorithm of Zelt and Smith (1992) was used to interpret the<br />

data. The model was constructed with 13 boundaries, each of which contained up to 38 boundary nodes to<br />

8


cover a distance of 530 km. These boundaries form 12 layers, each specified by up to 14 velocity nodes<br />

for the top and bottom of the layer. Model parameters are linearly interpolated between <strong>adjacent</strong> nodes<br />

during ray tracing and display. All OBSs were deployed along a line, except OBS 13. It is slightly off line<br />

and in the subsurface <strong>adjacent</strong> to a salt dome in the Orpheus Graben. OBS 13 recorded the worst quality<br />

data.<br />

We use “a, b, c and d” to label crustal boundaries, and Pa, Pb, Pc, and Pd for P-wave <strong>refraction</strong>s<br />

from underneath them. Similarly, Sa, Sb, Sc, and Sd represent S-wave <strong>refraction</strong>s. PmP and SmS<br />

represent P- and S-wave <strong>wide</strong>-<strong>angle</strong> <strong>reflection</strong>s from the Moho. The vertical resolution general decreases<br />

with depth as the wavelengths increase (0.2-1.0 km). The lateral resolution in the upper crust in this study<br />

is determined primarily by the spacing of the OBS. The scale over which lateral variations are averaged in<br />

the upper crust is about half the receiver spacing or ~20 km (White 1989). Below the upper crust the<br />

resolution is limited by the ray path geometries associated with the arrivals modelled from these layers<br />

(Zelt and Forsyth 1994). Their generic results for the middle and lower crust are that velocities are not<br />

well constrained because most of the ray paths are reflected phases. In this study the middle crustal<br />

velocity of 6.4 km/s is constrained by refracted arrivals. However, in the lower crust and upper mantle,<br />

poor lateral resolution is caused by the refracted waves restricted to near-horizontal ray paths. An error<br />

analyses is performed, by perturbing model parameters until the computed arrival times no longer provide<br />

an adequate fit to the data on one or more OBS sections. In general, the error bounds from several OBS<br />

are smaller than or equal to that for a single OBS section. For example, the velocity uncertainty is<br />

sometimes found to be 0.15 km/s for a single OBS, but when several OBS are considered, it reduces to<br />

0.10 km/s.<br />

9


Sedimentary Layers<br />

Along <strong>profile</strong> 99-1, the lowest velocity layer is associated with the 70-km-<strong>wide</strong>,


instrument. On OBS 6 and 7, the velocity of the first arrivals is greater than 6.0 km/s. Eastward, modelling<br />

of the near offset arrivals on OBS 8, 9, 10 and 11 indicate that the velocities range from 5.9-6.1 km/s,<br />

below a velocity of about 5.5 km/s that is only 0.5 km thick.<br />

Crust and Moho<br />

For the upper crust, the observed amplitudes are usually strong on each OBS, modelled by a<br />

velocity of 6.0 km/s that increases downward to produce diving waves in the upper 5-km-thick crust (Fig.<br />

5). The amplitudes decrease dramatically beyond 50 km offset. This is modelled with a velocity of 6.0-6.3<br />

km/s and a low gradient in the upper 20 km depth. As an example, OBS 1 (Fig. 5) records clear P- and S-<br />

waves at


this layer the velocity structure is not directly constrained by our data, although OBS 12 and 14 do record<br />

deeper crustal arrivals as diving waves from underneath the graben.<br />

The crustal layers below “b” and “c” have a low vertical velocity gradient. All the data from OBS<br />

1-8 exhibit a relatively low velocity of 6.4-6.6 km/s between boundary “c” and the Moho. As a result,<br />

only upper crustal <strong>refraction</strong>s and PmP can be clearly observed. In comparison, <strong>reflection</strong> <strong>profile</strong> 89-2 that<br />

crosses the <strong>refraction</strong> line near OBS 5 (Fig. 4) clearly shows a transparent upper crust, somewhat<br />

reflective middle crust, and more reflective lower crust. This highly reflective lower crust is also exhibited<br />

on many OBS in the Meguma Terrane (e.g. OBS 6 and 8; Figs. 6 and 7).<br />

The lack of turning rays in the Meguma Terrane crust makes the <strong>wide</strong>-<strong>angle</strong> <strong>reflection</strong> PmP<br />

clearly visible in nearly all our OBS <strong>profile</strong>s in this terrane (Fig. 8). The PmP <strong>reflection</strong> is not a single<br />

<strong>reflection</strong> event but occurs over an interval of at least a second. The lower crustal velocity is primarily<br />

determined based on the curvature of PmP and rare intra-crustal <strong>reflection</strong>s. To determine the sensitivity<br />

of the seismic arrivals to lower crustal velocity, the velocities were perturbed by 0.2 km/s increments.<br />

The effect of this change is exemplified on OBS 6 (Meguma) and 10 (Avalon) shown in Figure 9. For<br />

OBS 6, the computed travel-time curve of PmP with the correct velocity of 6.6 km/s matches the<br />

curvature of observed amplitudes, whereas when this velocity is perturbed to 6.8 km/s, the computed<br />

curvature does not match. On the contrary, PmP on OBS 11 has to be modelled by a higher velocity of<br />

6.9 km/s. It can be seen that a velocity of 6.6 km/s does not generate the correct curvature (Fig. 9b).<br />

Therefore, the resolution of the lower crustal velocities is estimated to be better 0.2 km/s. This observed<br />

crustal velocity in Meguma Terrane is lower than in cratons but is consistent with the average Paleozoic<br />

crust (Holbrook et al. 1992).<br />

12


In the crust from the west end of the model to the Orpheus Graben, the velocity gradient is low.<br />

However <strong>reflection</strong>s are recorded on most of the OBS (Fig .10) that indicate velocity discontinuities,<br />

labelled as “a”, “b”, “c” and “d” do occur. In the lower crust on OBS 9, 10 and 12 (Fig. 5) stronger<br />

<strong>reflection</strong>s occur that are associated with a more significant velocity discontinuity. We also marked as<br />

solid segments on the model (Fig. 10) the range of the actual reflecting points from the Moho.<br />

OBS 14 (Fig. 11), located on the Avalon Terrane, recorded PmP and SmS, and Pn up to 350 km<br />

offset to the west. Modelling of the PmP of this OBS indicates a shallower (34 km depth) Moho<br />

underneath the Orpheus Graben. The Moho deepens westward to 38 km depth according to OBS 12, and<br />

shallows to 37 km depth further west. Pn is observed on OBS 6, 11 and 14, and its associated velocity<br />

from the uppermost mantle is estimated at 7.950.1 km/s. The picked arrivals for the complete data set<br />

overlain with computed travel time curves from the final model are shown in Figure 12.<br />

Shear waves<br />

Shear waves (or S-waves) refracted from upper and lower crust are observed on many OBS, such<br />

as OBS 1 (Fig. 5a, b), 8, and 14 (Fig. 11b). The picking of their arrivals is usually more difficult than for<br />

corresponding P-waves due to reverberations of the earlier P-waves. Iterative modelling suggests that we<br />

can obtain a general fit using a constant Vp/Vs ratio for each layer in the velocity model for Vp. Poisson’s<br />

ratios as derived from modelling the S-waves are shown in Figure 10 and 11. For the upper 10 to 15 km<br />

(laterally variable) the ratio is 0.21. It can be seen that, for the mid and lower crustal layers, the Poisson’s<br />

ratio is = 0.24. Barrett et al. (1964) determined values of Poisson’s ratio of 0.22 for entire crust of the<br />

Meguma Terrane consistent with ratios calculated here.<br />

13


Although this is a simplified description of the crust, computed travel-time curves of SmS, or the<br />

<strong>wide</strong>-<strong>angle</strong> S-wave <strong>reflection</strong> from the Moho, fit the observed data reasonably well. An example is OBS<br />

14 (Fig. 11b); the observed SmS is composed of many seconds of reverberations with its first arrival<br />

much weaker than later arrivals. After coherency mixing of the record section, the first arrival time of<br />

SmS can be picked near 145-170 km distance, while the general curvature of the SmS is clearly defined,<br />

and therefore modelled, from its multiple reverberations.<br />

Laboratory Velocity and Density Measurements<br />

To provide a quantitative basis for interpreting the seismic results discussed above in terms of<br />

petrology, the acoustic velocities of a large suite of rock samples from <strong>Nova</strong> Scotia that could be<br />

considered candidates for the middle and lower crust under the Meguma Terrane were measured in the<br />

laboratory at elevated confining pressures ranging from zero to 1 GPa to simulate in situ conditions. The<br />

samples included granulite facies rocks from the lower crustal Tangier xenolith suite (Giles and<br />

Chatterjee, 1987), high pressure rocks from the Liscomb Complex (Clarke et al. 1993), granites and<br />

tonalities from the Musquodoboit batholith and coastal and offshore plutons, plus samples of the Tangier<br />

dykes lamprophyres and the Meguma formation itself.<br />

The laboratory measurements were made using the pulse transmission technique of Birch (1960)<br />

and Christensen (1985) on 2.54 cm diameter, 3 to 6 cm long minicores drilled from the samples and<br />

trimmed to right cylinders. After the densities of the samples were determined using the immersion<br />

method, the minicores were jacketed with thin copper foil to prevent the pressure medium from invading<br />

the samples at high pressures. One MHz lead zirconate piezoelectric transducers and backup electrodes<br />

were placed on the ends of the minicores and the entire assembly was jacketed in gum rubber tubing. The<br />

14


samples were then placed in a hydrostatic pressure vessel at the GSC/ Dalhousie University High Pressure<br />

Laboratory in Halifax, N.S., Canada and compressional wave velocities (Vp) were measured at selected<br />

pressure intervals by sending a high voltage spike to the transmitting transducer and calculating the<br />

velocity from the length of the minicore and the time of flight of the signal between transducers. Identical<br />

procedures were used to measure shear wave velocities (Vs) except that lead zirconate titanate transducers<br />

with matching polarities were used instead of lead zirconate transducers. The velocities shown in Table I<br />

were recorded as the pressure decreased and are considered accurate to 0.5% for Vp and 1.0% for Vs<br />

(Christensen and Shaw, 1970), while the densities are considered accurate to + 0.005 g/cm 3 . As noted by<br />

many investigators, the measured velocities increase rapidly with pressure to about 200 MPa due to the<br />

closure of microcracks, but at higher pressures, the velocities tend to increase slowly and linearly with<br />

pressure in response to the intrinsic elastic properties of the constituent minerals of the rocks themselves.<br />

As can be seen in Figure 13, in which Vp and Vs are plotted versus density at 600 MPa, the in situ<br />

pressure at a mid-crustal depth of about 20 km, the granites and tonalites cluster fairly tightly about a<br />

mean density of 2.67 g/cm 3 and mean compressional and shear wave velocities of 6.18 and 3.59 km/s,<br />

respectively, while the mafic lithologies display a mean density of 2.88 g/cm 3 and mean compressional<br />

and shear wave velocities of 6.72 and 3.77 km/s, respectively. Similarly, at 1 GPa, the pressure at the<br />

base of the crust, the granites and tonalities display average compressional and shear wave velocities of<br />

6.25 and 3.62 km/s, respectively, while the mafic rocks average 6.76 and 3.80 km/s. Interestingly, the<br />

high pressure Liscomb and Tangier xenolith assemblages both display a much <strong>wide</strong>r range in velocity,<br />

with Vp ranging from about 6.8 to 8.0 km/s at 600 MPa, with the highest velocity samples being<br />

metapelites rich in sillimanite and sapphirine.<br />

15


Gravity model<br />

To model the gravity anomaly a simple two-dimensional algorithm was used. We extract a<br />

<strong>profile</strong> from a dense regional data grid (Fig. 3)( Dehler and Roest 1998), interpolated along the <strong>refraction</strong><br />

<strong>profile</strong> at an interval of about 1 km. The final velocity model in Figure 10 is divided into rectangular<br />

blocks, each bordered by boundary or velocity nodes of the velocity model. Each block is assumed to<br />

have a constant density , calculated from the average velocity (v) of the block, using an empirical<br />

relationship: = -0.6997 + 2.2302 v -0.598 v 2 + 0.07036 v 3 -0.0028311 v 4 . A maximum density of 3.30<br />

10 3 kg m -3 is assigned to the upper mantle.<br />

Although gravity modelling does not generate unique results, it can provide a valuable check on<br />

our velocity model, especially the Moho topography. For example, the modelled depth of the shallower<br />

Moho underneath the Orpheus Graben is sensitive to the longer wavelength of the computed gravity,<br />

whereas structures closer to the seabed (e.g. inside the graben) influence the shorter wavelengths of the<br />

computed gravity curves. In the area near the Orpheus Graben, the seismic methods have the weakest<br />

control. Therefore, we combined both methods to constrain the model. It can be seen that the shape of the<br />

observed and calculated gravity anomalies are similar (Fig. 10), with the greatest mismatch in the Orpheus<br />

Graben. This mismatch may be attributed to the oblique crossing of the WAR line with the graben, and to<br />

the lack of detailed control inside the graben. Gravity modelling of the Orpheus Gravity anomaly by<br />

Loncarevic et al. (1967) indicated a feature with a depth of 10 km similar to that determined by the<br />

seismic experiment.<br />

16


Discussion<br />

The P-wave velocity of 5.5-5.8 km/s observed from the southwest end of the line to the Orpheus<br />

Graben (Fig. 10) is attributed to the metasedimentary Meguma Supergroup. This is deduced from the<br />

character of the magnetic anomalies supported by offshore geological samples and based on comparison<br />

with laboratory velocity measurements (Table I). The S-wave velocity of 3.2 km/s is lower than the range<br />

of velocities measured on the Meguma Supergroup samples 3.52-3.8 km/s (Table I). This may be due to<br />

heterogeneities of the lithologies within the Supergroup and the small number of laboratory<br />

measurements.<br />

The velocity model (Fig. 10) shows a maximum thickness of 3 km of this unit at the southwest<br />

segment of the line. Our velocity and thickness are similar to the range of measurements of Barrett et al.<br />

(1964) on this section of the line. Gravity modelling of the SMB (Benn et al. 1999) suggests that it was<br />

emplaced at the base of the Meguma Supergroup limiting the Supergroup’s thickness to 7 km. It is<br />

probable that uplift and erosion has significantly thinned this unit. Thus, we interpret the Meguma<br />

Supergroup as a thin sheet and suggest it is now a roof pendant.<br />

Orpheus Graben on the Cobequid-Chedabucto Fault and Avalon Terrane<br />

The most striking features of the near surface arrivals along the WAR <strong>profile</strong> are the delays in the<br />

first arrivals due to the thick low velocity section of 2.5-2.8 km/s and 4.4-5.6 km/s in the Orpheus Graben<br />

attributed to sedimentary rocks (Fig. 5b). The WAR velocities are constrained by well velocities for the<br />

upper 3 km and the gravity modelling suggest that the graben is a maximum of 10 km thick. The crustal<br />

velocities beneath the graben and to the east show evidence of being higher throughout the crust (Fig. 10).<br />

Similar velocities for the lower crust are recorded for the Avalon Terrane off Newfoundland (Hall et al.<br />

17


1998). The Avalon Terrane is measured to be 40 km deep on WAR line 91/2 along the Atlantic coast of<br />

Newfoundland and about 12 s or greater based on <strong>reflection</strong> <strong>profile</strong> 84-2 northeast of Newfoundland and<br />

offshore of southern New Brunswick (Keen et al. 1991a).<br />

The surface expression of the CCF (Webster et al., 1998; Pe-Piper and Piper 2004) is plotted on<br />

Figure 3. Its sub-surface shape at the northeastern end our study is based on <strong>reflection</strong> line 86-5 (Fig. 10).<br />

This <strong>reflection</strong> <strong>profile</strong> shows an inclined plane, similar to that required by our <strong>refraction</strong> data but<br />

perpendicular to it. The fault shape, crustal thickness and velocities suggest that Avalon Terrane underlies<br />

the Meguma Terrane as far southwest as OBS 7. Along <strong>profile</strong> 99-2 another control point for the extent of<br />

the Avalon Terrane is available. In the Bay of Fundy delimitation is based on <strong>reflection</strong> <strong>profile</strong>s 88-4 and<br />

88-2. We prefer the interpretation of Keen et al. (1991a) with the Avalon zone terminating near the<br />

beginning of <strong>profile</strong> 88-2 (Fig. 3). There is a crustal ramp on the <strong>reflection</strong> <strong>profile</strong> 88-4 that surfaces in the<br />

vicinity of the CCF (Withjack et al, 1995, Pe-Piper and Piper 2004) and tapers southward. The depth to<br />

Moho is 12s beneath the Avalon Terrane. The crustal wedge thins and terminates to the southward on 88-<br />

2 where a highly reflective lower crust and flat Moho at 10s are interpreted as the Meguma Terrane.<br />

Based on these control points from seismic data and extrapolating between them, the subsurface<br />

extent of the Avalon Terrane beneath the Meguma is also plotted on the gravity map (Fig. 3). These points<br />

connect parallel to the CCF. It is striking that the deep feather-edge of the Avalon Terrane parallels the<br />

negative gravity anomaly associated with the SMB and its offshore extent and terminates prior to the<br />

SMB. Geochemical data support this interpretation. Major element analysis of D-26 (Fig. 1) (Pe-Piper and<br />

Jansa 1999) falls within the average of the SMB analysis. In contrast, the samples from the F-22 well have<br />

Nd and Pb isotopes compositions similar to the Avalon Terrane (Pe-Piper and Jansa 1999).<br />

18


Crustal layer Meguma Terrane<br />

From the southern end of the <strong>profile</strong> 99-1 to the Orpheus Graben along the Meguma Terrane, the<br />

P-wave velocity of the crust increases from 6.0 km/s near the surface to 6.6 km/s above the Moho. This<br />

velocity range is compared to the laboratory measurement (Table I, Fig. 13). Assuming average heat flow<br />

values similar to those in the eastern United States (Blackwell, 1971; Lachenbruch and Sass, 1977), T will<br />

be about 300 o C at 20 km depth and 600 o C at the base of the crust. Assuming further that dVp/dT = -.39<br />

and -.55 x 10 -3 km/s o C, respectively, for granitoid and mafic rocks (Christensen and Mooney, 1995) and<br />

dVs/dT = -.15 and -.21 x 10 -3 km/s o C for the same lithologies (after Ji et al, 2002), Vp will be ~6.06 and<br />

6.01 km/s and Vs~3.53 and 3.50 km/s for granitoid rocks in the middle and lower crust; and Vp will be<br />

~6.57 and 6.46 km/s for mafic rocks in the middle and lower crust, while Vs will be ~3.69 and 3.65 km/s.<br />

Poisson’s ratio, which is relatively insensitive to P and T, since Vp and Vs tend to rise and fall together, is<br />

~ 0.24 for the granitoids and ~0.27 for the mafics, suggesting that granitoids dominate the crust and that<br />

mafics are only significant near the base of the section, perhaps interlayered with granitic rocks to cause<br />

the observed reflectivity. Another possibility if the temperature is slightly elevated is that the higher P-<br />

wave and Poisson’s ratio values near the base of the section could be caused by a phase change from to<br />

quartz which occurs at about 650 o C (Barruol 1993).<br />

Seismic <strong>reflection</strong> <strong>profile</strong> 89-2 (Fig. 4) shows a transparent upper crust to a depth of about 15 km<br />

compatible with a thick batholith (Fig. 4). Gravity modelling of the SMB indicates a thickness of 6 km<br />

(Benn et al. 1999). If for the batholith we estimate an average thickness of 10 km, a length of 500 and a<br />

width of 100 km, then the volume of granitoid rocks is 500,000 km 3 . There must be equally voluminous<br />

ultramafic restites below the batholith (eg Wolf and Wyllie 1994). One possible explanation for the lack<br />

19


of mafic crust is that it foundered into the mantle. Another is that the seismic and petrologic Mohos are<br />

not the same. Mafic-ultramafic rocks in the lower crust could be difficult to distinguish from the<br />

peridotites of the mantle. Further consideration of the origin of the crust is discussed later in this section.<br />

Development of the crustal structure<br />

We consider various hypotheses that can explain the inverted stratigraphy of the Meguma Terrane<br />

(Eberz et al., 1991, Clark et al 1993). One way for the older metasedimentary rocks to overly a younger<br />

crust is a tectonic break. Seismic <strong>reflection</strong> line 88-1 exhibits a spectacular group of dipping reflectors that<br />

could be consistent with thrusting (Keen et al. 1991b). However, the depth of these <strong>reflection</strong>s is too great<br />

to be consistent with the 3 km thick section of metasedimentary rocks overlying the younger granitic<br />

crust. Furthermore, a tectonic thrust does not explain the even level exposure of the crust, and the lack of<br />

exposure of lower crustal rocks and ophiolites.<br />

A hot spot model has also been used to describe the voluminous short-lived magmatism of the<br />

northern Appalachians (Murphy et al. 1999). This model is not in agreement with the felsic nature of the<br />

crust. Furthermore, a hot spot would be expected to have underplated the crust with a mafic layer with<br />

velocities in the range of 7.0-7.6 km/s. We found no evidence of such a layer.<br />

Another explanation for the voluminous granitic rocks is regional extension (Pe-Piper and Jansa<br />

1999). The timing of the granitoid and mafic intrusions is late to post orogenic. The plutonism continued<br />

into the Carboniferous on the Scotian margin when the area was undergoing regional extension. Other<br />

evidence for extension is the <strong>wide</strong> across strike extent of the plutons, the ubiquitous extension in the<br />

Magdalen Basin, and the geochemical and isotopic variation of the granites that are attributed to<br />

20


decompression melting. Indeed, the elongated shape of the SMB is parallel to the CCF. However the other<br />

granite intrusives were emplaced along the northeast trending folds of the Meguma Supergroup (eg<br />

Keppie and Dallmeyer 1995) perpendicular to the CCF arguing against across strike extension to solely<br />

account for the emplacement of the granitoid bodies. Other evidence that extension does not account<br />

entirely for the formation of the granites are folding and shearing foliations in two batholiths that recorded<br />

tectonic deformation as they crystallized, indicating syn-tectonic emplacement (Benn et al. 1997).<br />

A more inclusive premise for the development of the Meguma Terrane that includes the<br />

generation of the granitoid rocks, the young source ages for the lower crust and the uplift history is<br />

detachment of the lower crust and upper lithosphere. Delamination (Nelson 1992) is the rapid mechanical<br />

thinning of the mantle beneath an orogenic belt (Fig. 14). Delamination has been previously proposed for<br />

the Meguma Terrane (Keppie and Dallmeyer 1995) and in the Gander zone in Newfoundland (Schofield<br />

and D’Lemos 2000). The sequence of events begins with a collision that thickens the crust and<br />

lithosphere, followed by detachment of the lower crust and lithosphere that initiates uplift and erosion<br />

followed by extension and subsidence.<br />

The Meguma Supergroup’s sedimentology is consistent with its formation in a volcanic arc<br />

setting. There are locally occurring volcaniclastic rocks and bulk chemical analysis suggests a strong<br />

volcanic component. In addition, the overlying Annapolis Supergroup is formed of subaerial<br />

volcaniclastic rocks (Schenk 1997). The meta-igneous xenoliths of Tangier dyke major and trace element<br />

compositions and ratios (Eberz et al. 1991), and the composition matrix (Giles and Chatterjee 1987) are<br />

consistent with crustal thickening in a continental volcanic arc. Greenough et al. (1989) based on U-Pb<br />

zircon and monazite dates for the xenoliths suggest that shortening between the Meguma and Avalon<br />

terranes occurred ~ 400 Ma and emplacement of the xenolith bearing dykes at ~370 Ma.<br />

21


The granulite-facies xenoliths found in the Tangier dyke are a feature diagnostic of delaminated<br />

terranes (Costa and Rey 1995). The P/ T conditions of these xenoliths suggest they come from greater<br />

depths than the present Moho (Giles and Chatterjee 1987, pressures of 12-14 kbars and temperatures of<br />

>1000 0 C). Based on the metamorphic grade of the xenoliths, Keppie and Dallmeyer (1995) advocate that<br />

the crust was 65 km thick. The subsequent foundering of the lower crust would explain the discrepancy<br />

with the present Moho depth.<br />

Delamination of the entire crustal root could result in complete melting of the lower crust (Nelson<br />

1992). Any region that has extensive granites such as observed in the Meguma Terrane may be inferred to<br />

have had temperatures that are hot enough for the lower crust to flow. Keppie and Dallymeyer (1995)<br />

propose that the emplacement of the granites about 30 Ma after initial convergence is consistent with the<br />

slow rise of the geotherms as the crust and lithosphere were thickened. They also hypothesize that the<br />

oldest pluton at 380-378 Ma was the time of the detachment of the lower crust. The composition of the<br />

metasedimentary xenoliths are consistent with the SMB being derived from them (Eberz et al. 1991).<br />

Seismic signatures of delamination are a crustal thickness that is thinner than the <strong>adjacent</strong> craton<br />

and a highly reflective lower crust (Nelson 1992) (Fig. 10 and 14). The consequence of detachment of the<br />

crustal root for the overlying rocks is uplift that is observed in the Meguma Terrane. Geobarometric data<br />

indicate that the present surface of the batholiths were emplaced at 6-10 km depth (Raeside and Mahoney<br />

1997). Based on a regional unconformity (Martel et al. 1993), within 13-23 million years of their<br />

emplacement the granitoid plutons were exhumed. The rate of erosion is consistent with ductile thinning<br />

of the crust and post-orogenic collapse. Exhumation erosion rates are < 5 km Ma -1 , in contrast to the faster<br />

exhumation due to synorogenic erosion and normal faulting. Following uplift a thermal sag basin is<br />

created in this model. This is compatible with the once <strong>wide</strong> spread distribution of 4-5 km Late<br />

22


Carboniferous to Early Permian strata over the Meguma Terrane (Ryan and Zentilli 1993). In summary,<br />

the delamination is one hypothesis that is consistent with many of the characteristic observed in the<br />

Meguma Terrane and explains its inverted stratigraphy.<br />

Conclusions<br />

A velocity layer of 5.5-5.7 km/s, attributed to the Meguma Supergroup, is 3 km thick or less<br />

<strong>adjacent</strong> the coast. The thinness of this section may be due to erosion. The velocity of the crust is 6.0-6.6<br />

km/s. Poisson’s ratio for the crust ranges from 0.21 to 0.24 indicates it is felsic. The upper portions of the<br />

crust are transparent on seismic <strong>reflection</strong> <strong>profile</strong>s. Granitoid rocks that form 25 % of the surface<br />

exposure on shore have both the appropriate P- and S -wave velocities and low reflectivity that make them<br />

a prime candidate for the upper crust. In the Meguma Terrane reflectivity is particularly high in the lower<br />

crust. From both the Liscomb Complex and the Tangier xenolith suite, there are high and low velocity<br />

rocks that could cause this. The suture between the Avalon and Meguma terranes is constrained to a<br />

southward ramp of 60 km in length. The SMB lies outboard and parallel to the edge of the subsurface<br />

extent of the Avalon Terrane. Four models for the tectonic history of the Meguma Terrane were<br />

considered. Components of several of the models are incorporated in a detachment hypothesis that is<br />

consistent with our velocity/depth model.<br />

Acknowledgments<br />

We would like to thank all the people who participated in the field experiment both on shore and<br />

at sea. We are particularly grateful to the Master, Captain Marsden, Officers and Crew of the CGGS<br />

Hudson whose excellent seamanship made launching and recovery of the airgun array and the ocean<br />

23


ottom seismometers possible in a variety of sea states. The Master and Crew of the forty foot CGGS<br />

Frank M Westin who launched the OBS on line 1 to save ship time for the larger vessel are thanked for<br />

their enthusiastic assistance. We thank Dr. S. Dehler who provided the plot files of the potential field data<br />

and B. Iuliucci who ran the samples through the high pressure laboratory. Dr. Clarke, Dr. S Dehler, and<br />

Dr. C. Hawkins provided comments that have substantially improved the manuscript.<br />

24


References<br />

Barrett, D.L., Berry, M., Blanchard, J.E., Keen, M.J. and McAlister, R.E. 1964. Seismic studies on the<br />

eastern seaboard of Canada: the Atlantic coast of <strong>Nova</strong> Scotia. Canadian Journal of Earth<br />

Sciences, 1: 10-22.<br />

Barruol, G., 1993. Petrophysique de la croute inferieure. Role de l’anisotropie sismique sur la reflectivite<br />

et le dephasage des ondes. PhD Thesis Universite de Montpellier II, France, pp. 239-258.<br />

Benn, K., and Pignotta, G.S. 1999. Magnetic fabric of the Barrington Passage pluton, Meguma Terrane,<br />

<strong>Nova</strong> Scotia, a two stage fabric history of syntectonic emplacement, Tectonophysics,107: 75-92.<br />

Benn, K., Jorne, R.J., Kontak, D.J., Pignotta, G.S. and Evans, N.G. 1997. Syn-Acadian emplacement<br />

model for the South Mountain batholith, Meguma Terrane, <strong>Nova</strong> Scotia: Magnetic fabric and<br />

structural analyses Geological Society of America Bulletin, 109: 1279-1293.<br />

Benn, K., Roest, W.R., Rochette, P., Evans, N.G. and Pignotta, G.S. 1999. Geophysical and structural<br />

signatures of a syntectonic batholith construction: the South Mountain Batholith, Meguma<br />

Terrane <strong>Nova</strong> Scotia. Geophysical Journal International, 136: 144-158.<br />

Birch, F., 1960. The velocity of compressional waves in rocks to 10 kilobars, 1, Journal of Geophysical<br />

Research, 65: 1083-1102.<br />

Blackwell, D.D., 1971. The thermal structure of the continental crust, in The Structure and Physical<br />

Properties of the Earth’s Crust, Geophysical Monograph Series, vol. 4, edited by J.G. Heacock,<br />

pp. 196-184, AGU, Washington, D.C.<br />

25


Christensen, N.I. 1985. Measurements of the dynamic properties of rocks at elevated temperatures and<br />

pressures, American Society for Testing and Materials, Special technical Publication 869:9 3-107.<br />

Christensen, N.I. 1989. Reflectivity and the seismic properties of the lower crust. Journal of Geophysical<br />

Research, 93: 1087-1102.<br />

Christensen, N.I., and Shaw, G.H., 1970. Elasticity of mafic rocks from the mid-Atlantic Ridge,<br />

Geophysical Journal of the Royal Astronomical Society, 20: 271-284.<br />

Christensen, N.I., and Mooney, W.D. 1995. Seismic velocity structure and composition of the continental<br />

crust: A global view, Journal of Geophysical Research, 100: 9761-9788.<br />

Clarke, D.B., and Chatterjee, A.K. 1992. Origin of peraluminous granites in the Meguma Zone of<br />

southern <strong>Nova</strong> Scotia: a synthesis. Joint Annual Meeting of GAC/AGC, Wolfville, abstract, 17:<br />

A18.<br />

Clarke, D.B., Halliday, A.N., and Hamilton, P.J. 1988. Neodymium and strontium isotopes constraints on<br />

the origin of the peraluminous granitoids of the South Mountain Batholith, <strong>Nova</strong> Scotia. Chemical<br />

Geology, 73: 15-24.<br />

Clarke, D.B., Chatterjee, A.K., and Giles, P.S. 1993. Petrochemistry, tectonic history, and Sr-Nd<br />

systematics of the Liscomb Complex, Meguma Lithotectonic Zone, <strong>Nova</strong> Scotia. Canadian<br />

Journal of Earth Sciences, 30: 449-464.<br />

26


Clarke, D.B. , Fallon. R., amd Heamann, L.M. 2000. Interaction among upper crustal , lower crustal and<br />

Matle material in the Port Mouton pluton, Meguma Lithotectonic Zone, southwest, <strong>Nova</strong> Scotia,<br />

Canadian Journal of Earth Sciences, 37: 576-600.<br />

Costa, S., and Rey, P. 1995. Lower crustal rejuvenation and growth during post-collisional collapse:<br />

insights from a core crust section through a Variscan metamorphic core complex. Geology, 23:<br />

905-908.<br />

Dehler, S.A., and Roest, W.R. 1998. Gravity Anomaly map, Atlantic Region. Geological Survey of<br />

Canada Open File 3658, scale 1: 3,000,000.<br />

Eberz, G.W., Clarke, D.B., Chatterjee, A.K., and Giles, P.S. 1991. Chemical and isotopic composition of<br />

the lower crust beneath the Meguma Lithotectonic Zone, <strong>Nova</strong> Scotia: evidence from granulite<br />

facies xenoliths. Mineralogy and Petrology, 109: 69-88.<br />

Giles, P.S., and Chatterjee, A.K. 1987. Lower crustal xenocrysts and xenoliths in the Tangier dyke,<br />

eastern Meguma Zone, <strong>Nova</strong> Scotia, <strong>Nova</strong> Scotia Department of Mines and Energy Report of<br />

Activities, Part A 87-5: 85-88.<br />

Greenough, J.D., Ruffman, A., and Owen, J.V. 1988. Magma injection directions inferred from a fabric<br />

study of the Popes Harbour dike, eastern shore, <strong>Nova</strong> Scotia, Canada, Geology, 16: 547-550.<br />

Greenough, J.D., Krogh, T.E., Kamo, S.L., Owen, J.V., and Ruffman, A. 1999. Precise U-Pb dating of the<br />

Meguma basement xenoliths: new evidence for Avalonian underthrusting. Canadian Journal of<br />

Earth Sciences, 36: 15-22.<br />

27


Hall, J., Marillier, F., and Dehler, S. 1998. Geophysical studies of the structure of the Appalachian orogen<br />

in the Atlantic borderlands of Canada, Canadian Journal of Earth Sciences, 35: 1205-1221.<br />

Ham, L.J. 1998. Musquodoboit Batholith Project: geological mapping of the Musquodoboit Harbour area<br />

(NTS 11D/14) <strong>Nova</strong> Scotia Department of Natural Resources, Report of Activities, 15-19.<br />

Holbrook, W.S., Mooney, W.D., and Christensen, N.I. 1992. The seismic velocity of the deep continental<br />

crust, Chapter 1 in Continental Lower crust edited by D.M. Fountain, R. Arculus, and R. W. Kay,<br />

Developments in Geotectonic, 23: 1-43.<br />

Jackson, H.R., Chian, D., Salisbury, M. and Shimeld, J. 2000 Preliminary crustal structure from the<br />

Scotian margin to the Maritimes Basin from <strong>wide</strong>-<strong>angle</strong> <strong>reflection</strong>/<strong>refraction</strong> <strong>profile</strong>s, Geocanada<br />

2000 The Millennium Geoscience Conference Cdrom, May 29-June 2, Calgary, Alberta<br />

Jansa, L. 1991. Regional Geology and Geophysics 9: Pre-Mesozoic basement and Mesozoic igneous rock<br />

occurrences: In East Coast Basin Atlas Series: Scotian Shelf, Atlantic Geoscience Centre,<br />

Geological Survey of Canada, p.25.<br />

Ji, S., Wang, Q., and Xia, B., 2002. Handbook of Seismic Properties of Minerals, Rocks and Ores, Ecole<br />

Polytechnique de Montreal, pp. 630.<br />

Keen, C.E., Kay, W.A., Keppie, D., Marillier F., Pe-Piper, G., and Waldron, J. 1991a. Deep seismic<br />

<strong>reflection</strong> data from the Bay of Fundy and the Gulf of St. Lawrence: tectonic implications for the<br />

northern Appalachians. Canadian Journal of Earth Sciences, 28: 1096-1111.<br />

28


Keen, C.E., MacLean, B.C., and Kay, W.A. 1991b. A deep seismic <strong>reflection</strong> <strong>profile</strong> across the <strong>Nova</strong><br />

Scotian continental margin, offshore eastern Canada. Canadian Journal of Earth Sciences, 28:<br />

1112-1120.<br />

Keen, C.E., and Potter, D.P. 1995. Formation and evolution of the <strong>Nova</strong> Scotian rifted margin: Evidence<br />

from deep seismic <strong>reflection</strong> data. Tectonics, 14: 918-932.<br />

Keppie, J.D., and Dallmeyer, R.D. 1987. Dating transcurrent accretion: an example from the Meguma and<br />

Avalon composite terranes in the northern Appalachians. Tectonics, 6: 831-847.<br />

Keppie, J.D., and Dallmeyer, R.D. 1995. Late Paleozoic collision, delamination , short-lived magmatism,<br />

and rapid denudation in the Meguma Terrane, (<strong>Nova</strong> Scotia, Canada): constraints from 40 Ar/ 39 Ar<br />

isotopic data. Canadian Journal of Earth Sciences, 32: 644-659.<br />

Lachenbruch A.H., and Sass, J.H. 1977. Heat flow and the thermal regime of the crust. In The Earth’s<br />

Crust: Its Nature and Physical Properties, Geophysical Monograph Series, vol. 20, Edited by J.G.<br />

Heacock, pp. 626-675, AGU, Washington, D.C.<br />

Loncarevic, B.D. 1967. Geophysical study of the Orpheus gravity anomaly. Proceedings VII World<br />

Petroleum Conference, pp. 828-835.<br />

Loncarevic, B.D., Courtney, R.C., Fader, G.B.J., Giles, P.S., Piper. D.J.W., Costell, G., Hughes Clark,<br />

J.E., and Stea, R.R. 1994. Sonography of a glaciated continental shelf. Geology, 22: 747-750.<br />

29


Marillier, F., Keen, C.E., Stockmal, G.S., Quinlan, G., Williams, H., Colman-Sadd., S.P., and O’Brian,<br />

S.J. 1989. Crustal structure and surface zonation of the Canadian Appalachians: implications of<br />

deep seismic <strong>reflection</strong> data. Canadian Journal of Earth Sciences, 26: 305-321.<br />

Martel, A.J., McGregor, D.C., and Utting, J. 1993. Stratigraphic significance of Upper Devonian and<br />

Lower Carboniferous miospores from the type area of the Horton Group, <strong>Nova</strong> Scotia. Canadian<br />

Journal of Earth Sciences, 30: 1091-1098.<br />

Murphy, J.B., van Staal, C.R. and Keppie, J.D.1999. Middle to Late Paleozoic Acadian orogeny in the<br />

northern Appalachians: a Larmide-style plume-modified orogeny? Geology, 27: 653- 656.<br />

Nelson, K.D. 1992. Are crustal thickness variations in old mountain belts like the Appalachians a<br />

consequence of lithospheric delamination? Geology, 20: 498-502.<br />

Oakey, G.N., and Dehler, S.A. 1998. Magnetic Anomaly map, Atlantic Region. Geological Survey of<br />

Canada Open File 3659, scale 1: 3,000,000.<br />

Pe-Piper, G., and Jansa, L.F. 1999. Pre-Mesozoic basement rocks offshore <strong>Nova</strong> Scotia, Canada: New<br />

constraints on the accretion history of the Meguma terrane. Geological Association of America<br />

Bulletin, 111: 1773-2792.<br />

Pe-Piper,G., and Loncarevic, B.D. 1989. Offshore continuation of the Meguma Terrane, southwest <strong>Nova</strong><br />

Scotia. Canadian Journal of Earth Sciences, 26: 176-191.<br />

30


Pe-Piper, G., and Piper, D. J.W. 2004. The effects of strike-slip motion along the Cobequid-Chedabucto<br />

Fault southwest Grand banks fault system on the Cretaceous-Tertiary evolution of Atlantic<br />

Canada. Canadian Journal of Earth Sciences, 41: 799-808.<br />

Raeside, R.P., and Mahoney, K.M. 1997. The contact metamorphic aureole of the South Mountain<br />

Batholith, southern <strong>Nova</strong> Scotia, Geological Association of Canada Annual Meeting Abstract, 21<br />

A-77.<br />

Ryan, R. J., and Zentilli, M. 1993. Allocyclic and thermochronological constraints on the origin of the<br />

Maritimes Basin of eastern Canada. Atlantic Geology, 29: 187-197.<br />

Schenk, P. E. 1970. Regional variation of the flysch-like Meguma Group (Lower Paleozoic) of <strong>Nova</strong><br />

Scotia, compared to recent sedimentation off the Scotian shelf. The Geological Association of<br />

Canada Special Paper, 7: 127-153.<br />

Schenk, P. E. 1971. Southeastern Atlantic Canada, northwestern Africa, and continental drift. Canadian<br />

Journal of Earth Sciences, 8: 1218-125.<br />

Schenk, P.E. 1995. Meguma Zone; In Chapter 3 of the Geology of the Appalachian - Calodonian Orogen.<br />

Edited by H. Williams; Geological Survey of Canada, Geology of Canada, no. 6: 261-277 (also<br />

Geological Society of America, The Geology of North America, v. F-1).<br />

Schenk, P.E. 1997. Sequence stratigraphy and provenance on the Gondwana’s margin: The Meguma Zone<br />

(Cambrian to Devonian) of <strong>Nova</strong> Scotia, Canada. Geological Society of America Bulletin, 109:<br />

395-409.<br />

31


Schofield, D.I., and D’Lemos, R.S. 2000. Granite petrogenesis in the Gander Zone, NE Newfoundland:<br />

mixing of melts from multiple sources and the role of lithospheric delamination. Canadian Journal<br />

of Earth Sciences, 37: 525- 547.<br />

Wade, J.1991. Scotian shelf lithostratigraphy, Pre-Mesozoic basement, Mohican formation and<br />

equivalent. In East Coast Basin Atlas Series: Scotian Shelf, Atlantic Geoscience Centre,<br />

Geological Survey of Canada, p. 55.<br />

Webster, T.L., Murphy, J.B. and Barr, S.M. 1998. Anatomy of a terrane boundary, an integrated<br />

structural, geographic information system and remote sensing study of the Late Paleozoic Avalon-<br />

Meguma terrane boundary drift. Canadian Journal of Earth Sciences, 35: 787-801.<br />

White, D.J. 1989. Two dimensional seismic <strong>refraction</strong> ray tomography. Geophysical Journal International,<br />

97: 223-247.<br />

Williams, H., and Hatcher, R. D. 1982. Suspect terranes and accretionary history of the Appalachian<br />

Orogen. Geology, 10: 530-536.<br />

Withjack, M.O., Olsen, P.E., and Schlische, R.W. 1995. Tectonic evolution of the Fundy rift basin,<br />

Canada: Evidence of extension and shortening during passive margin development. Tectonics, 14:<br />

390-405.<br />

Wolf, M.B. and Wyllie, P.J. 1994. Dehydration-melting of amphibolite at 10 kbar: the effects of<br />

temperature and time. Contributions to Minerology and Petrology, 115: 369-388.<br />

32


Zelt, C.A., and Smith, R.B. 1992. Seismic travel time inversion for 2-D crustal velocity structure.<br />

Geophysical Journal International, 108: 16-34.<br />

Zelt, C.A., and Forsyth, D.A. 1994. Modelling <strong>wide</strong> <strong>angle</strong> seismic data for crustal structure: Southeastern<br />

Grenville Province. Journal of Geophysical Research, 99: 11,687-11,704.<br />

33


Figures<br />

Figure 1) Location of <strong>wide</strong>-<strong>angle</strong> <strong>reflection</strong> /<strong>refraction</strong> <strong>profile</strong>s 99-1 and 2 relative to the onshore geology<br />

of <strong>Nova</strong> Scotia. The Cobequid-Chedabucto Fault (CCF) is shown separating the Meguma and Avalon<br />

terranes. SMB and MB in italics indicate the South Mountain and Musquodoboit batholiths, the sample<br />

locations at various rock types are indicated by: T for the Tangier dyke, L for the Liscomb Complex, M<br />

for the Musquodoboit Batholith, Me for the Meguma Supergroup. The dashed line 64-1 is the position of<br />

the <strong>refraction</strong> line of Barrett et al. (1964) and the thin solid line are deep multichannel <strong>reflection</strong> <strong>profile</strong>s.<br />

Figure 2) Geological time chart summarizing the evolution of the Meguma Terrane.<br />

Figure 3) Free Air (offshore) and Bouguer gravity (onshore) anomaly map. Hollow circles with labels<br />

indicate the location of wells and short bedrock cores. The crosses show the position of the Ocean Bottom<br />

Seismometers along the WAR <strong>profile</strong>. Deep multichannel seismic <strong>profile</strong>s in the region are shown as<br />

solid white lines. The solid black lines and dashed to dotted white lines represent respectively the surface<br />

and subsurface expression of the Cobequid-Chedabucto Fault.<br />

Figure 4) Multichannel seismic <strong>reflection</strong> <strong>profile</strong> 89-2 that crosses the WAR <strong>profile</strong>s 99-1 and 2. Strong<br />

coherency mixing was done to highlight the characteristics of the Meguma Terrane crust. Transparent<br />

upper crust (UC), reflective lower crust (LC) are labelled. The boundaries a, b, d and Moho are based on<br />

the WAR velocity model where it crosses the <strong>reflection</strong> <strong>profile</strong>. The seismic <strong>reflection</strong> <strong>profile</strong> was<br />

converted to depth using our velocity model.<br />

34


Figure 5) Computed travel time curves are overlain on the record sections of OBS 1and 12. The horizontal<br />

scale is shot-receiver distance (offset) and the vertical scale is the travel time using a reduction velocity of<br />

6.5 km/s for the P-wave. In 5b the P and S refracted waves from the metasedimentary layer and the upper<br />

crust above boundary a are shown. P and S wave velocities in km/s are shown as 5.5 (3.3). is Poison’s<br />

ratio followed by the P to S ratio in brackets. 5c OBS 12 at the edge of the Orpheus Graben shows the<br />

contrast in arrival patterns with OBS 1 located on the Meguma Terrane. 5d Refl LC indicates the zone of<br />

reflective lower crust.<br />

Figure 6) P-wave (a) and S-wave (c) record sections of OBS 6, showing typical features of the WAR<br />

Profiles along 99-1. The horizontal scale in the record section is shot-receiver distance (offset) and the<br />

vertical scale is the travel time using a reduction velocity of 6.5 km/s. The small vertical gradient below<br />

boundary a in (b) causes the Pa and Sa to terminate abruptly at ~ 60 k m offset. Abundant <strong>wide</strong>-<strong>angle</strong><br />

<strong>reflection</strong>s are produced between boundary d and the Moho modelled by lenses of high and low velocities.<br />

Section (b) illustrates the few arrivals between boundaries a and d. The numbers are velocities are in km/s.<br />

Computed travel time curves are overlain on the data in (a and c).<br />

Figure 7a) OBS 8 record section is overlain with the travel time curves. The horizontal scale in the record<br />

section is shot-receiver distance (offset) and the vertical scale is the travel time using a reduction velocity<br />

of 6.5 km/s. 7b is a plot of the ray tracing.<br />

Figure 8) Sections of OBS 6, 7, 9, 10 and 11 emphasizing the lower crust and Moho <strong>reflection</strong>s. The large<br />

numbers at the upper edge of the record sections are the OBS numbers. All seismic <strong>profile</strong>s have the same<br />

35


horizontal and vertical scales while the velocity model is horizontally compressed. LC: lower crust. The<br />

dotted lines with arrows indicate the position of the record section on the <strong>profile</strong>s.<br />

Figure 9) The effects of different velocities in the lower crust on the computed travel time curves plotted<br />

on the record sections from OBS 6 and 11. The horizontal scale is shot-receiver distance (offset) and the<br />

vertical scale is the travel time using a reduction velocity of 6.5 km/s. The solid and dashed curves are<br />

computed PmP arrivals at various velocities.<br />

Figure 10) Calculated (solid line) and observed free air gravity anomalies are plotted across the top. The<br />

velocities are the largest numbers in km/s. Poisson’s ratio is shown as . The numbers in the ellipses<br />

indicate the particular OBS that constrain the layer at that point. The arrows on the top of the <strong>profile</strong><br />

indicate where there is a <strong>reflection</strong> <strong>profile</strong> that crosses the WAR <strong>profile</strong>. The P-36 is the name of the well<br />

that is used to control the velocities in the Orpheus Graben The bottom section of the panel shows the<br />

seismic <strong>reflection</strong> <strong>profile</strong>s that are available at an <strong>angle</strong> to the WAR <strong>profile</strong>s to constrain the <strong>wide</strong> <strong>angle</strong><br />

data.<br />

Figure 11) OBS 14 P- and S-wave arrivals (a) and (c) with the model for both shown in (b).<br />

Figure 12) Comparison of observed and calculated travel times for OBS 1 to 14, shown with the<br />

corresponding ray paths. The observed data are indicated by vertical bars with their heights representing<br />

the uncertainty of the picks; calculated data are indicated as solid lines. Vertical scale is in seconds with a<br />

reduction velocity of 6.5 km/s applied to the travel times and horizontal scale is offset in kilometers. The<br />

number of the OBS is indicated on the upper left corner of each section. OBS 13 was excluded because it<br />

was off line.<br />

36


Figure 13. Compressional (Vp) and Shear wave (Vs) velocity versus density for <strong>Nova</strong> Scotia samples at<br />

600 MPa. G, granitoid rocks; M, mafic rocks. Line L-T shows range of velocities for Liscomb and<br />

Tangier xenolith assemblages. Also shown are lines of constant acoustic impedance, Z; an impedance<br />

difference of 2.5 is sufficient to give a strong seismic <strong>reflection</strong>.<br />

Figure 14) A cartoon of the development of the Meguma Terrane in the context of a delamination model.<br />

The steps are: 1) the formation of the Meguma Terrane, 2) collison, 3) delamination, 4) extension and 5)<br />

collapse.<br />

37


TABLE I. LABORATORY VELOCITIES AND DENSITIES OF NOVA SCOTIA ROCK SUITES<br />

Sample Rock Type Density Mode Velocity (km/s) at Pressure (MPa)<br />

Meguma<br />

38<br />

g/cc 100 200 400 600 800 1000<br />

GO-1 greywacke 2.71 P 5.89 5.94 6.01 6.06 6.10 6.14<br />

2.71 S 3.71 3.74 3.77 3.78 3.79 3.80<br />

Hib-1 silicic greywacke 2.75 P 6.04 6.11 6.14 6.16 -- --<br />

CH-1 staur-sill schist 2.84 P 6.01 6.09 6.16 6.20 -- --<br />

HFX-3 slate 2.83 P 6.42 6.49 6.58 6.63 -- --<br />

Musquodoboit batholith<br />

2.83 S 3.52 3.55 3.59 3.61 -- --<br />

MI-1 granite 2.64 P 6.21 6.30 6.39 6.44 6.49 6.53<br />

Coastal and Offshore plutons<br />

2.64 S 3.66 3.71 3.75 3.76 3.77 3.77<br />

NS-17 granite 2.70 P 5.81 5.91 6.00 6.05 -- --<br />

NS-20 granite 2.67 P 5.48 5.60 5.72 5.78 -- --<br />

2.67 S 3.19 3.23 3.28 3.31 -- --


NS-21 granite 2.64 P 5.99 6.11 6.20 6.25 -- --<br />

39<br />

2.64 S 3.53 3.59 3.62 3.65 -- --<br />

NS-Y17 tonalite 2.74 P 5.99 6.11 6.20 6.25 -- --<br />

2.74 S 3.36 3.43 3.51 3.56 -- --<br />

NS-Y18 tonalite 2.66 P 5.67 5.88 6.00 6.08 -- --<br />

NS-Y20A tonalite 2.72 P 6.00 6.11 6.22 6.25 -- --<br />

2.72 S 3.55 3.59 3.64 3.67 -- --<br />

NS-420B tonalite 2.71 P 6.36 6.48 6.61 6.69 -- --<br />

Liscomb Complex<br />

LG-6 granite 2.62 P 5.79 5.94 6.06 6.12 6.16 6.17<br />

2.62 S 3.52 3.57 3.62 3.65 3.68 3.70<br />

LG-30 granite 2.63 P 5.91 6.02 6.10 6.14 6.18 6.21<br />

2.63 S 3.61 3.68 3.70 3.70 3.71 3.71<br />

LG-189 granite 2.63 P 5.99 6.03 6.15 6.23 6.28 6.30<br />

2.63 S 3.31 3.37 3.40 3.43 3.45 3.46<br />

LG-130-3 gneiss 2.84 P 5.97 6.10 6.21 6.25 6.27 6.31<br />

LG-98 augen gneiss 2.65 P 5.93 6.10 6.26 6.33 6.36 6.39<br />

2.65 S 3.60 3.65 3.69 3.71 3.73 3.75<br />

BI-98-1 gabbro 2.89 P 6.15 6.33 6.46 6.48 6.49 6.50<br />

2.89 S 3.27 3.32 3.37 3.38 3.39 3.39


LG-133-1 gneiss 2.82 P 6.27 6.36 6.45 6.50 6.52 6.54<br />

40<br />

2.82 S 3.66 3.70 3.75 3.77 3.78 3.78<br />

LG-178-1 gabbro 2.85 P 6.27 6.44 6.56 6.60 6.61 6.62<br />

2.85 S 3.52 3.60 3.68 3.71 3.72 3.72<br />

LG-172 gabbro-diorite 2.92 P 6.48 6.59 6.65 6.67 6.68 6.69<br />

2.92 S 3.67 3.72 3.76 3.78 3.79 3.80<br />

LG-146 gabbro 2.86 P 6.59 6.69 6.79 6.83 6.84 6.85<br />

2.86 S 3.76 3.78 3.80 3.81 3.83 3.84<br />

LG-165 gabbro 2.81 P 6.58 6.68 6.77 6.82 6.85 6.87<br />

2.81 S 3.63 3.68 3.72 3.74 3.76 3.77<br />

LG-162-2A mafic cumulate 2.93 P 6.67 6.84 7.00 7.10 7.15 7.19<br />

2.93 S 3.48 3.56 3.65 3.70 3.72 3.74<br />

LG-120 sill schist 2.97 P 7.53 7.60 7.68 7.74 7.81 7.87<br />

Tangier dyke lamprophyres<br />

2.97 S 4.13 4.18 4.20 4.23 4.26 4.28<br />

LG-197-1 lamprophyre 2.80 P 6.10 6.19 6.30 6.37 6.41 6.44<br />

2.80 S 3.47 3.52 3.57 3.60 3.61 3.61<br />

LG-194 lamprophyre 2.84 P 6.28 6.36 6.45 6.50 6.54 6.58<br />

Tangier dyke xenoliths<br />

2.84 S 3.61 3.64 3.66 3.67 3.69 3.70


LG-192x-23 metasediment 2.90 P 5.61 5.70 5.79 5.85 5.90 5.94<br />

sa+gt+sp>50% 2.90 S 3.36 3.46 3.53 3.57 3.59 3.59<br />

LG-192x-25 metasediment 2.71 P 5.66 5.79 5.97 6.04 6.08 6.10<br />

cpx+gt>50% 2.71 S 3.38 3.41 3.45 3.47 3.48 3.49<br />

LG-195x-24 metasediment 2.91 P 5.67 5.95 6.13 6.18 6.21 6.23<br />

opx+cpx>50% 2.91 S 3.31 3.38 3.47 3.52 3.55 3.57<br />

LG-192x-1 metasediment 2.82 P 5.97 6.06 6.14 6.20 6.25 6.29<br />

opx+gt>50% 2.82 S 3.59 3.65 3.69 3.71 3.72 3.73<br />

LG-192x-10 metasediment 2.82 P 5.84 6.01 6.17 6.22 6.36 6.30<br />

opx+gt>50% 2.82 S 3.57 3.58 3.60 3.62 3.63 3.64<br />

LG-196x-23 metabasalt 2.81 P 6.05 6.21 6.32 6.38 6.42 6.44<br />

amph>50% 2.81 S 3.68 3.73 3.79 3.83 3.85 3.86<br />

192x-6 metasediment 2.80 P 6.27 6.32 6.36 6.39 6.42 6.45<br />

sa+gt+sp>50% 2.80 S 3.41 3.48 3.55 3.59 3.62 3.65<br />

LG-192x-17 metasediment 2.77 P 6.17 6.25 6.34 6.40 6.43 6.46<br />

cpx+gt>50% 2.77 S 3.83 3.87 3.90 3.91 3.92 3.92<br />

192x-8 metasediment 2.88 P 6.31 6.35 6.38 6.41 6.45 6.48<br />

cpx+gt>50% 2.88 S 3.80 3.81 3.83 3.84 3.85 3.85<br />

TD-98-1 metasediment 2.77 P 6.13 6.27 6.40 6.44 6.47 6.50<br />

cpx+gt>50% 2.77 S 3.69 3.73 3.75 3.75 3.76 3.76<br />

TD-92-3 metasediment 2.82 P 6.38 6.44 6.50 6.55 6.57 6.59<br />

cpx+gt>50% 2.82 S 3.63 3.65 3.68 3.70 3.71 3.71<br />

41


196x-23 metabasalt 2.79 P 6.19 6.41 6.52 6.56 6.60 6.63<br />

amph>50% 2.79 S 3.54 3.61 3.68 3.72 3.75 3.76<br />

LG-192x-22 metabasalt 2.83 P 6.38 6.48 6.57 6.62 6.66 6.68<br />

amph>50% 2.83 S 3.49 3.56 3.64 3.66 3.69 3.69<br />

196x-26A metabasalt 2.80 P 6.28 6.46 6.58 6.64 6.68 6.71<br />

amph>50% 2.80 S 3.54 3.62 3.71 3.75 3.79 3.80<br />

192x-7 metasediment 2.86 P 6.64 6.70 6.76 6.81 6.84 6.87<br />

opx+gt+sp>50% 2.86 S 3.47 3.49 3.50 3.52 3.53 3.54<br />

LG-197x-1 metabasalt 3.01 P 6.55 6.77 6.95 7.03 7.06 7.07<br />

cpx>50% 3.01 S 3.81 3.87 3.92 3.96 3.98 3.99<br />

196x26 metabasalt 3.05 P 6.39 6.60 6.83 6.95 7.02 7.08<br />

cpx>50% 3.05 S 3.94 4.06 4.16 4.19 4.19 4.20<br />

LG-192x-4 metasediment 3.00 P 6.86 6.95 7.05 7.11 7.16 7.20<br />

sa+gt+sp>50% 3.00 S 3.81 3.84 3.88 3.90 3.91 3.92<br />

192x-3 metasediment 2.96 P 7.76 7.86 7.94 7.98 8.02 8.07<br />

sa+gt+sp>50% 2.96 S 3.88 3.95 4.02 4.04 4.06 4.07<br />

abbreviations: silic, silicified; st, staurolite; and, andalusite; sill, sillimanite; gt, garnet; hb, hornblende; sa, sappharine; sp,<br />

spinel; opx, orthopyroxene; cpx, clinopyroxene; amph, amphibole; P, compressional wave; S, shear wave.<br />

42


88-2<br />

88-4<br />

CCF Bay of Fundy<br />

Carboniferous and Younger<br />

Sedimentary and Volcanic rocks<br />

Liscomb Complex<br />

Devonian Intrusive Rocks (Granites)<br />

Early Paleozoic Metasediments<br />

and Volcanics (Meguma Supergroup)<br />

North Mountain Basalt<br />

Rock Sampled for Vp-Vs<br />

Shelburne Dyke<br />

Recorders for 64-1 Line<br />

OBS Location for 99-1 and 99-2<br />

SMB<br />

MEGUMA AVALON<br />

64-1<br />

99-2<br />

Me<br />

be Gr<br />

CCF O pheus<br />

L<br />

M<br />

89-2 50 0 m<br />

PEI<br />

T<br />

64-2<br />

99-1<br />

1 2 3 4 6 7 8 9 10 11<br />

0 km<br />

MB<br />

99-2b<br />

100<br />

88-1<br />

Cp ae Breton<br />

r a n<br />

N ORTH AMERICA 12<br />

ATLANTIC<br />

OCEAN<br />

Fig. 1


TABLE OF FORMATIONS<br />

AGE<br />

Ma<br />

ERA PERIOD OROGENY<br />

66<br />

144<br />

208<br />

245<br />

286<br />

320<br />

360<br />

374<br />

387<br />

408<br />

438<br />

505<br />

570<br />

1000<br />

3000<br />

PROTEROZOIC PALEOZOIC MESOZOIC CZ<br />

QUATERNARY<br />

CRETACEOUS<br />

JURASSIC<br />

TRIASSIC<br />

PERMIAN<br />

CARBONI-<br />

FEROUS<br />

DEVONIAN<br />

SILURIAN<br />

ORDOVICIAN TACONIAN<br />

CAMBRIAN<br />

HADRYNIAN<br />

HELIKIAN<br />

HERCYNIAN<br />

deformation<br />

ACADIAN<br />

AVALONIAN<br />

GRENVILLIAN<br />

EVOLUTION OF MEGUMA TERRANE<br />

final rifting of Atlantic<br />

underplating SW Scotia Margin<br />

Stretching of Pangea; Shelburne dyke<br />

diabase; N. Mountain Basalt<br />

4 km erosion of Maritimes Basin<br />

subsidence<br />

formation of Maritimes Basin<br />

exhumation of granite<br />

intrusion of peraluminious granites,<br />

mafic dykes<br />

accretion of Meguma<br />

to N. America<br />

Meguma Group deposition<br />

formation of components of<br />

Meguma sediments:<br />

(Sm/Nd systematics and zircons)<br />

Fig. 2


67°W 65°W 63°W 61°W 59°W 57°W<br />

48°N<br />

88-2<br />

88-4 0<br />

km<br />

100<br />

Core 22<br />

99-1 B-93<br />

I-94<br />

99-2<br />

86-1<br />

PEI<br />

89-2 8<br />

1 6 7<br />

N-30<br />

E-7<br />

88-1<br />

CCCCFF 12<br />

P-36<br />

I-22<br />

865 -<br />

14<br />

D-26<br />

E-53<br />

F-38<br />

F-52<br />

50 30 10 -10 -30 -50 -70<br />

mGal<br />

89-1<br />

Fig. 3<br />

46°N<br />

44°N


Depth (km)<br />

0<br />

5<br />

10<br />

15<br />

20<br />

25<br />

30<br />

35<br />

40<br />

45<br />

SW NE<br />

MCS 89-2<br />

Refraction 99-1/99-2<br />

0 Distance (km) 20<br />

UC<br />

LC<br />

Moho<br />

a<br />

b<br />

d<br />

Fig. 4


PmP<br />

T-X/6.5 (s)<br />

12<br />

T-X/6.5(s)<br />

Depth (km)<br />

5<br />

4<br />

2<br />

1<br />

0<br />

0<br />

4<br />

W<br />

OBS 1<br />

(a)<br />

(d)<br />

(b)<br />

a<br />

0 km 10<br />

OBS 12<br />

(c)<br />

a<br />

b<br />

c<br />

d<br />

6.4<br />

6.3<br />

5.5(3.3)<br />

5.8(3.4)<br />

= .24(1.69)<br />

6.0(3.6)<br />

.05<br />

= .21(1.65)<br />

6.1(3.7)<br />

6.2(3.8)<br />

= .21(1.65)<br />

Distance (km)<br />

40<br />

Refl LC<br />

Moho<br />

6.5<br />

6.6 6.7-6.9<br />

6.9 8.0<br />

7.9<br />

0 50 100 150 200 250 300<br />

0 km 50<br />

Depth (km) -145-130 -115 -100 -85 -70<br />

Distance (km)<br />

OBS 2<br />

d<br />

a<br />

Sa<br />

Pa<br />

E<br />

Fig. 5


T-X/6.5(s)<br />

Depth(km)<br />

T-X/4.0(s)<br />

W E<br />

7<br />

6<br />

4<br />

2<br />

0<br />

40<br />

-150<br />

12<br />

10<br />

8<br />

6<br />

4<br />

2<br />

OBS 6 P-Wave<br />

(a)<br />

Pn<br />

(b)<br />

0<br />

-150<br />

a<br />

b<br />

c<br />

d<br />

-100<br />

OBS 6 S-Wave<br />

(c)<br />

-100<br />

6.2<br />

5.9<br />

6.5<br />

6.4<br />

-50<br />

-50<br />

6.3<br />

6.6<br />

PmP d<br />

SmS<br />

0<br />

50<br />

Distance (km)<br />

d<br />

a<br />

0<br />

50<br />

Distance (km)<br />

S2<br />

100<br />

100<br />

a<br />

6.7-6.9<br />

a<br />

150<br />

150<br />

Pn<br />

Fig. 6<br />

210<br />

Comlexity<br />

Near<br />

Orpheus<br />

Graben<br />

210


Depth (km)<br />

T-X/6.5 (s) 7 WE<br />

0<br />

OBS 8 d<br />

(a)<br />

d<br />

(b)<br />

a<br />

b<br />

c<br />

d<br />

Moho<br />

6.3<br />

6.4<br />

Refl LC<br />

6.5<br />

b<br />

6.6<br />

PmP<br />

40<br />

-100 -50<br />

0 50 100 150<br />

200 250<br />

Distance (km)<br />

a<br />

6.7<br />

6.9<br />

Fig. 7


Time-X/6.5 (s)<br />

Depth (km)<br />

6<br />

4<br />

PmP<br />

Refl<br />

LC<br />

2<br />

6<br />

4<br />

2<br />

W E<br />

6<br />

6<br />

4<br />

0<br />

10<br />

20<br />

30<br />

40<br />

10<br />

Avalon LC<br />

7<br />

2<br />

-120 -30<br />

a<br />

b<br />

c<br />

d<br />

Refl LC<br />

6.4<br />

6.3<br />

2<br />

-100 -40<br />

7<br />

m<br />

d<br />

Avalon LC<br />

7<br />

m<br />

2<br />

OBS 6 7 9 10 11<br />

6.5<br />

6.6<br />

?<br />

130 200 240<br />

0 90<br />

Avalon LC<br />

-150 -100 -50 0 50 100 150 200 250 300 350<br />

9<br />

9<br />

10<br />

7<br />

11<br />

0 50 100 20 60 100 120<br />

Distance (km)<br />

6.7<br />

6.9<br />

m2 ?<br />

Pn<br />

m<br />

6.7<br />

6.9<br />

Fig. 8


6<br />

4<br />

2<br />

W 6<br />

E<br />

Time - X/6.5 (sec)<br />

OBS 11<br />

5<br />

4<br />

3<br />

OBS 6<br />

Pn<br />

-100 -50<br />

Lower Crustal Velocity for Meguma Terrane:<br />

V=6.6<br />

6.8?<br />

20 50 100 120<br />

Distance (km)<br />

Lower Crustal Velocity for Avalon Terrane:<br />

V=6.9<br />

6.6?<br />

PmP<br />

PmP Fig. 9<br />

PmP (V= 6. 6?)


86-5b 89-2 99-1 2<br />

99- REFR99-1<br />

-100 mGal 100<br />

E<br />

MCS86-5b<br />

Avalon14<br />

4.2-4.8<br />

O. Graben<br />

P-36<br />

12<br />

11<br />

8<br />

7<br />

6<br />

MCS89-2<br />

REFR99-2<br />

5<br />

(Halifax)<br />

4<br />

5.5-5.8<br />

4.3<br />

5.9<br />

5.9<br />

6.5?<br />

5.5-5.8<br />

OBS 1 2<br />

.21(1.65)<br />

a<br />

1<br />

.21(1.65)<br />

W<br />

0<br />

.24<br />

1.69<br />

5<br />

Salts<br />

3<br />

6.0-6.1 .05<br />

6.2 .10<br />

6.1<br />

Meguma<br />

9 10<br />

.05<br />

10<br />

10<br />

6.2<br />

6.4<br />

6.3<br />

1<br />

60 .<br />

6.<br />

11<br />

9<br />

7<br />

8<br />

6<br />

6.2?<br />

4<br />

5<br />

0<br />

6<br />

5.<br />

9<br />

8<br />

.22<br />

<br />

b<br />

10<br />

6.7?<br />

6.8?<br />

Meguma Terrane<br />

2<br />

8<br />

.24(1.71)<br />

15<br />

10<br />

6.3<br />

6.4<br />

c<br />

20<br />

Marilliar et al.<br />

1994<br />

.24(1.71)<br />

Depth (km)<br />

2<br />

1<br />

9<br />

Avalon Terrane<br />

6.4<br />

8<br />

25<br />

5<br />

.1<br />

68 .<br />

8<br />

4<br />

6.5<br />

6<br />

10<br />

d<br />

30<br />

9<br />

14<br />

14<br />

.1<br />

10<br />

11 6.9<br />

7.95 .1<br />

6 11 14<br />

6<br />

5<br />

.1<br />

6.6<br />

3 4<br />

7<br />

6<br />

5<br />

.1<br />

.24(1.71)<br />

6.6<br />

35<br />

6.9?<br />

12<br />

9<br />

7<br />

2<br />

40<br />

0 50 100 150 200 250 300 350<br />

45<br />

-150 -100 -50<br />

N<br />

Avalon<br />

REFR<br />

99-1<br />

O. Graben<br />

Meguma<br />

REFR<br />

99-1/2<br />

SW NE S<br />

0<br />

0<br />

5<br />

UC<br />

10<br />

b<br />

10<br />

LC<br />

15<br />

20<br />

20<br />

25<br />

30<br />

30<br />

Depth (km)<br />

d<br />

MOHO<br />

40<br />

35<br />

40<br />

0 30 60 90<br />

45<br />

-150 -130 -110 -90<br />

MCS 86-5b<br />

MCS 89-2<br />

Fig. 10


T-X/7 (s)<br />

T-X/4.0(s)<br />

8<br />

6<br />

4<br />

2<br />

0<br />

45<br />

10<br />

8<br />

6<br />

4<br />

2<br />

0<br />

OBS 14 P-Wave<br />

(a)<br />

(b)<br />

a<br />

b<br />

c<br />

6.2(3.8)<br />

6.4<br />

=.21(1.65)<br />

d 6.5<br />

=.21(1.65)<br />

6.6<br />

OBS 14 S-Wave<br />

(c)<br />

SmS<br />

Pn<br />

8.0<br />

=.21(1.65)<br />

=.21(1. 65)<br />

63 .<br />

6.8<br />

0 50 100 150 200 250 300 350<br />

6.9<br />

d<br />

Distance (km)<br />

SmS d<br />

O Graben<br />

7.9<br />

PmP<br />

Fig .11


8<br />

6<br />

4<br />

2<br />

10<br />

20<br />

30<br />

40<br />

6<br />

4<br />

2<br />

10<br />

20<br />

30<br />

40<br />

8<br />

6<br />

4<br />

2<br />

10<br />

20<br />

30<br />

40<br />

6<br />

4<br />

2<br />

10<br />

20<br />

30<br />

40<br />

5<br />

3<br />

1<br />

-1<br />

10<br />

20<br />

30<br />

40<br />

1<br />

4<br />

7<br />

10<br />

14<br />

-100 0 100<br />

-100 0 100 200<br />

0 100 200 300<br />

2<br />

5<br />

-100 0 100 -100 0 100<br />

-100 0 100<br />

8 9<br />

11<br />

-100 0 100 200 -100 0 100 200<br />

3<br />

6<br />

12<br />

Fig .12


Velocity (km/sec)<br />

8<br />

7<br />

6<br />

5<br />

4<br />

3<br />

Vp<br />

Vs<br />

Lithology<br />

Meguma<br />

Musquodaboit, Plutons<br />

Liscomb granite<br />

Liscomb gneiss, schist<br />

Liscomb mafic<br />

Felsic Tangier xenoliths<br />

Mafic Tangier xenoliths<br />

Lamprophyre<br />

G<br />

G<br />

L-T<br />

M<br />

M<br />

Z= 17.5<br />

600MPa<br />

22.5<br />

2.5 3.0<br />

Density (g/cm3)<br />

3.5<br />

20<br />

Fig. 13


(1) 590-395 Ma<br />

(2) 395-388 Ma (3) ~ 355 Ma<br />

(4)<br />

SW<br />

Meguma deformation<br />

(5) 350-280 Ma<br />

35 km<br />

future<br />

Scotian margin<br />

Granites form<br />

385-370 Ma<br />

Meguma Supergroup Deposition<br />

CCF<br />

Meguma<br />

Avalon<br />

Meguma<br />

Meguma Avalon<br />

Carboniferous cover<br />

Meguma<br />

CCF CCF PP<br />

Avalon<br />

Delamination<br />

Moho<br />

150 km<br />

Exhumation of granites<br />

Moho<br />

40 km<br />

craton<br />

LEGEND<br />

CP – central pluton PP – peripheral pluton CCF– Cobequid Chedabucto Fault<br />

CP<br />

PP<br />

CCF<br />

Avalon<br />

Avalon<br />

NE<br />

Fig. 14

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!