Exploring the Deep Sea and Beyond themed issue

The natural range of submarine canyon-and-channel longitudinal profiles Jacob A. Covault*, Andrea Fildani, Brian W. Romans, and Tim McHargue Chevron Energy Technology Company, Clastic Stratigraphy Research and Development, 6001 Bollinger Canyon Road, San Ramon, California 94583, USA

ABSTRACT We differentiated 20 submarine canyonand-channel longitudinal profiles across various types of continental margins on the basis of relative convexity or concavity, and according to their similarities to best-fitting mathematical functions. Profiles are visually differentiated into convex, slightly concave, and very concave groups, each of which generally corresponds with a continental-margin type and distinct depositional architecture. Profile groups generally reflect the competing influences of uplift and construction of depositional relief of the seafloor and its degradation by erosion related to mass wasting. Longitudinal-profile shape provides a basis for classifying deep-sea sedimentary systems, linking them to the geomorphic processes that shape continental margins. INTRODUCTION Submarine canyon-and-channel systems are conduits through which sediment is transported across continental margins to deep-sea basins by sediment gravity flows and other mass movements (Shepard, 1948, 1981; Menard, 1955). Processes that sculpted canyon-and-channel systems during their lifetimes are manifested in the shapes of longitudinal profiles. Longitudinal profiles of fluvial systems have been contemplated by geomorphologists since the nineteenth century (e.g., Playfair, 1802; Gilbert, 1880). Subsequent studies of fluvial profiles have highlighted the relative importance of intrinsic and extrinsic controlling variables, including water discharge, sediment supply, sediment caliber (Snow and Slingerland, 1987), and changing uplift conditions. Interactions between these variables introduce complications into the characteristic logarithmic, or concave upward, shape of terrestrial fluvial profiles (Whipple *Present address: U.S. Geological Survey Eastern Energy Resources Science Center, 12201 Sunrise Valley Drive, Reston, Virginia 20192, USA.

and Tucker, 1999; Schumm et al., 2000). In contrast, previous work on submarine canyonand-channel profiles has been predominantly limited to case studies of profiles from a single type of continental margin (e.g., Goff, 2001; Estrada et al., 2005; Mitchell, 2005; Noda et al., 2008; Gerber et al., 2009). Tectonic and sedimentary influences inherent to different types of continental margins, however, have a significant impact on seascape morphology and sediment– gravity-flow erosion and deposition by controlling the gradient and stability of the seafloor, and sediment caliber and supply (Bouma et al., 1985; Normark, 1985; Stow et al., 1985; Mutti and Normark, 1987; Shanmugam and Moiola, 1988; Normark and Piper, 1991; Carvajal et al., 2009; Piper and Normark, 2009). Bill Normark pioneered work on seafloor canyon-and-channel systems and depositional fans in 1970 with “Growth Patterns of DeepSea Fans” (Normark, 1970). This seminal work prompted Normark to reconcile seafloor observations and interpretations of sediment–gravityflow processes with “ancient” buried subsurface and outcropping systems (e.g., 1982 COMFAN [COMmittee on FANs]; Bouma et al., 1985; Normark, 1985; Normark et al., 1985a; Mutti and Normark, 1987, 1991; Normark and Piper, 1991; Normark et al., 1993). These efforts produced broadly applicable models of sediment dispersal across continental margins that hold true to this day. Normark (1985) and Mutti and Normark (1987) highlighted that the characteristics of deep-water canyon-and-channel systems and depositional fans are controlled by the morphology and sediment supply characteristics of the submarine continental margin and receiving basin. Multiple continental-margin and receiving-basin scenarios and resultant deep-water seafloor and ancient stratigraphic architectures were recognized in an attempt to place a global and temporal breadth of turbidite architectures (i.e., seafloor, buried subsurface, and outcropping) within a common framework (Mutti and Normark, 1987). Following in the footsteps of Normark and colleagues’ seminal work on seafloor features

and implications for models of continental margins, this study reviews canyon-and-channel longitudinal profiles from a global range of continental margins—from the tectonically active transform and convergent margins of the Pacific to the Atlantic passive margin offshore the Americas (Fig. 1 and Table 1). We differentiated longitudinal profiles across continental slopes on the basis of (1) relative convexity or concavity from visual inspection and (2) according to their similarities to best-fitting mathematical functions. These observations and analyses provide a catalog of the breadth and general controls of the shapes of submarine sedimentdelivery systems, which can be related to the depositional architecture of different continental margins. Database and Methodology Our submarine canyon-and-channel database includes 20 longitudinal profiles from a variety of continental margins (Figs. 1 and 2; Table 1). These canyon-and-channel systems are submarine conduits that pass from predominantly erosional, V-shaped canyons indenting the shelf and uppermost slope to U-shaped channels with overbank deposits across the lower slope and continental rise (cf. Shepard, 1948; Menard, 1955; Normark, 1970). We examined profiles of canyons and channels that are present across the modern seafloor (i.e., buried channel features are excluded) and, as such, they represent the most recent canyon-and-channelsystem activity (i.e., since the last glacial cycle, <100 ka, for many systems; Lambeck and Chappell, 2001). Four canyon-and-channel profiles from the California Continental Borderland, La Jolla (14), Carlsbad (15), Oceanside (16), and Newport (17), were measured from National Oceanic and Atmospheric Administration/National Geophysical Data Center (NOAA/NGDC) and U.S. Geological Survey (USGS) multibeam bathymetry (<3 arc-second grids, 10-cm vertical resolution; Gardner and Dartnell, 2002; Dartnell et al., 2007; Divins and Metzger, 2009; Table 1). The Mississippi

Geosphere; April 2011; v. 7; no. 2; p. 313–332; doi: 10.1130/GES00610.1; 10 figures; 1 table.

For permission to copy, contact [email protected] © 2011 Geological Society of America

313

Covault et al.

19 20

7

5

18

9

17 16 15 Figure 1. Location map of canyon-and-channel systems analyzed in this study. Numbers correspond to systems in Table 1 and throughout the text.

8

13

6

1 3

14

10 12

2 11

4

profile (10) was measured from NOAA/NGDC multibeam bathymetry (Divins and Metzger, 2009; Table 1). We chose to measure the profiles of these five canyon-and-channel systems because (1) readily available high-resolution multibeam bathymetry covers their extents from the continental shelf out to the base of slope and (2) satisfactory published profiles were not discovered during the course of our literature review. The remaining profiles were compiled from published examples to facilitate further investigation of the canyons and channels of this study (Table 1). The majority of profiles were measured by other researchers from high-resolution multibeam bathymetric data (Table 1). The exact reference for each published canyon-and-channel longitudinal profile used in this study is emboldened and italicized in Table 1. Regardless of the resolution of published longitudinal profiles, long and short profiles are compared at similar resolutions as a result of the following measurement procedure: (1) water depths and down-system lengths were measured along every bend (cf. channel length and slope measurements of Flood and Damuth, 1987); and (2) water depths and down-system lengths of the high-resolution profiles were resampled every 10 km for systems >100 km long and every 1 km for systems <100 km long. Gradient and curvature (down-system change in gradient) were also calculated. Longitudinal-Profile Normalization Figure 3 shows all the longitudinal profiles used in this study. These are difficult to compare to one another because of differences in

314

canyon-and-channel lengths and depths. To rectify this problem, we have normalized longitudinal profiles in two ways: (1) on the basis of profile length from canyon head to the end of the confined portion of the system (e.g., at the channel-to-lobe transition zone of Mutti and Normark, 1987; Fig. 4); and (2) on the more objective basis of profile length from canyon head to the point where profiles reached gradients <0.25° and curvatures (i.e., down-system change in gradient) between –10–7 and 10–7. This second method focuses on the lengths of canyon-and-channel systems across their continental slopes rather than their lengths across relatively flat basin plains (cf. Adams and Schlager, 2000) (Fig. 5). Three problems arose when we attempted to normalize profiles across their predominantly confined segments (e.g., from canyon head to the channel-to-lobe transition): (1) not all of the profiles include the entire confined segments of their canyon-and-channel systems (i.e., many published examples and bathymetric data sets do not extend at sufficiently high resolution beyond the continental slope); (2) the exact locations of confined-to-unconfined transitions are commonly poorly defined, gradational sedimentary environments and, as a result, are subjective; and (3) even though profiles were from different continental margins, with very different tectonic and sedimentary controls on their development, they only fell into two groups of longitudinal profiles— convex upward and concave upward (Fig. 4A). Regarding problem 3, closer inspection of concave upward profiles shows that profiles from mature passive margins (e.g., Rhone [8], Zaire

Geosphere, April 2011

[11], and Amazon [12]) are only slightly concave in their proximal reaches (<0.4 of their total down-system length) relative to profiles from the margin offshore southern California, which are characterized by steeper slopes and narrower shelves (e.g., La Jolla [14], Carlsbad [15], Oceanside [16], Newport [17], and Ascension [19]) (Fig. 4B). All 20 of the profiles used in this study were normalized across their continental slopes (Fig. 5). This normalization procedure cuts off the relatively long and flat tail of a few of the longest profiles, which extend across basin plains (e.g., Mississippi [10], Zaire [11], and Amazon [12]) (Fig. 2). This is necessary in order to compare their reaches across slopes to other systems, which do not extend as far across the basin plain and are predominantly restricted to the slope. As mentioned above, the base of slope was not measured at an arbitrary or unit length; rather, it was defined at the point where profiles reached gradients <0.25° and curvatures between –10–7 and 10–7 (cf. Adams and Schlager, 2000) (Fig. 2). Here, we include all normalized longitudinal-profile data; however, we focus on normalization across the slope in order to provide a more inclusive comparison of profiles (Fig. 5). Longitudinal-Profile Groups and Curve-Fitting Functions Normalized longitudinal profiles were grouped in two ways: (1) on the basis of relative convexity or concavity from visual inspection; and (2) according to their similarities to bestfitting mathematical functions (cf. Shepherd,

Geosphere, April 2011

58

16

Normalized length across slope (km)

Fluvial sediment load (×106 t/yr) Mud to sand

40

76

Mud to gravel

N/A§

250

No

Yes

Linear

10 km3/yr**

Mud to pebbles

Mud to coarsegrained sand

200

Yes

Yes

Active, convergent

3.2 km3/yr#

110

No

Yes

Linear Active, convergent

SeaBeam and Hydrosweep bathymetry (measured over 5 km)

Hydrosweep bathymetry (better than 100-m horizontal resolution)

Active, convergent

Hokkaido, Japan, Pacific Ocean

Central Chile, Pacific Ocean

Exponential

200-m-resolution DEMs with a vertical accuracy of 0.6% of depth

Venezuela, Caribbean Sea

Presumably mud to pebbles

N/A§

200

No

Yes

Active, convergent

Linear

1980s-vintage SeaMARC II bathymetry (measured over <10 km)

Mud to gravel; silt rich

15

200

Yes

Yes

Active, convergen

Linear

1960s-vintage bathymetry (measured over <10 km)

Honshu, Japan, Washington, USA, Pacific Ocean Pacific Ocean

TABLE 1. CHARACTERISTICS OF CANYON-AND-CHANNEL SYSTEMS (3) Barbados A (4) San Antonio (5) Kushiro (6) Aoga (7) Astoria Convex Convex Convex Convex Slightly concave

1 Mud to gravel; mud rich

Mud to finegrained sand; silt rich

270

No†

No

Passive

Linear

7.4

170

Yes

No

Mixed*

Linear

SeaBeam bathymetry (measured over 5 km)

New Jersey, USA, Atlantic Ocean

France, Mediterranean Sea 1980s-vintage SeaBeam and SeaMARC I bathymetry

(9) Hudson Slightly concave

(8) Rhone Slightly concave

Mud rich

400

285

No

Yes; gravitydriven

Passive

Linear

NOAA/NGDC bathymetry (<3 arcsecond grids; 10-cm vertical resolution)

Gulf of Mexico

(10) Mississippi Slightly concave

References 21 15 12 2,20,29 2,23,25,29 5,17,26 17,25 16,18 2,29 3,4,17,25 Note: We measured five longitudinal profiles from high-resolution, multibeam bathymetric data sets that extend from the continental shelf out to the base of slope (Gardner and Dartnell, 2002; Dartnell et al., 2007; Divins and Metzger, 2009). The remaining profiles were compiled from published sources, which are highlighted in the row of references as emboldened and italicized numbers. See references list at the bottom of the table. (continued)

Mud to sand

No

Glacial influence (>40° latitude)

Grain size

Yes; gravitydriven

Yes; gravitydriven

Synsedimentary deformation No

Passive

Passive

Margin type

Best-fitting function Linear

3D seismicreflection data

3D seismicreflection data (processed bin size of 50 × 25 m)

Exponential

Niger Delta, Atlantic Ocean

NW Gulf of Mexico

Geographic context

Data

(2) Nigeria X Convex

(1) East Breaks Convex

System Longitudinal profile

Canyon-and-channel longitudinal profiles

315

316

Geosphere, April 2011

370

48

Normalized length across slope (km)

Fluvial sediment load (× 106 t/yr) 4 Mud to gravel; sand rich

Mud to pebbles; mud rich

360

Yes

No

Passive

1200

380

No

No

Passive

Logarithmic

1980s-vintage SeaBeam bathymetry

Eastern Canada, Atlantic Ocean

Mud to gravel; sand rich

Unknown, but likely small

2.2 (Oceanside littoral cell) Mud to gravel; sand rich

25

No

54

No

Yes

Mud to gravel; sand rich

0.4

58

No

Yes

Active, transform

Active, transform

Yes

Logarithmic

Logarithmic

NOAA/NGDC/ USGS bathymetry (<3 arcsecond grids; 10-cm vertical resolution)

Southern California, USA, Pacific Ocean

Active, transform

NOAA/NGDC/ USGS bathymetry (<3 arcsecond grids; 10-cm vertical resolution)

Southern California, USA, Pacific Ocean

Logarithmic

NOAA/NGDC/ USGS bathymetry (<3 arcsecond grids; 10-cm vertical resolution)

Southern California, USA, Pacific Ocean

Mud to gravel; sand rich

0.5

81

No

Yes

Active, transform

Logarithmic

NOAA/NGDC/ USGS bathymetry (<3 arcsecond grids; 10-cm vertical resolution)

Southern California, USA, Pacific Ocean

TABLE 1. CHARACTERISTICS OF CANYON-AND-CHANNEL SYSTEMS (continued) (13) Laurentian (14) La Jolla (15) Carlsbad (16) Oceanside (17) Newport Very concave Very concave Very concave Very concave Very concave

Sand rich

1.3

80

Mud to cobbles; sand rich

Unknown, but likely small

84

No

Yes; Qt terrace uplift, earthquakes common Yes

Yes

Mixed*

Logarithmic

NOAA bathymetric sheet NOS 1307N-11B

Central California, USA, Pacific Ocean

(19) Ascension Very concave

Active, transform

Logarithmic



France, Mediterranean Sea

(18) Var Very concave

Mud to cobbles; sand rich

4.8

140

No

Yes

Active, transform

Linear

Simrad bathymetry

Central California, USA, Pacific Ocean

(20) Monterey Slightly concave

References 1,29 2,9,24,25,29 2,17,28 6,10,13,22 6 6,10,14 10,14 27,29 8,11,19 2,11,29 *Mixed margins include settings adjacent to uplifting hinterland source areas (e.g., northwestern Mediterranean Sea). The northwestern Mediterranean Sea is a young, steep margin (post-Oligocene) with local tectonism (Savoye et al., 1993). † Hudson Canyon head is just south of 40° N latitude. § These canyons head on the middle continental slope, rather than the shelf edge. # Rio Miapo discharge (not fluvial sediment load). **Combined discharge of the Tokachi and Kushiro rivers (not fluvial sediment load). Note: References: (1) Babonneau et al. (2002); (2) Barnes and Normark (1985); (3) Beaubouef and Friedmann (2000); (4) Booth et al. (2000); (5) Butman et al. (2006); (6) Covault et al. (2007); (7) Curray et al. (2003); (8) Fildani and Normark (2004); (9) Flood et al. (1991); (10) Graham and Bachman (1983); (11) Greene et al. (2002); (12) Huyghe et al. (2004); (13) Inman (2008); (14) Inman and Jenkins (1999); (15) Klaus and Taylor (1991); (16) Laursen and Normark (2002); (17) Milliman and Syvitski (1992); (18) Milliman et al. (1995); (19) Nagel et al. (1986); (20) Nelson (1970); (21) Noda et al. (2008); (22) Normark (1970); (23) Normark et al. (1985b); (24) Pirmez and Imran (2003); (25) Pirmez et al. (2000); (26) Pratson et al. (1994); (27) Savoye et al. (1993); (28) Skene and Piper (2006); (29) Sømme et al. (2009). DEMs—digital elevation models; NOAA/NGDC—National Oceanic and Atmospheric Administration/National Geophysical Data Center; USGS—U.S. Geological Survey.

Mud rich

No

Glacial influence (>40° latitude)

Grain size

Yes; gravity-driven

Synsedimentary deformation

Passive

Margin type

Data

Linear

SeaBeam bathymetry (~100-m horizontal resolution)

Simrad bathymetry (better than 100-m horizontal resolution; 10-m vertical resolution)

Linear

Brazil, Atlantic Ocean

West Africa, Atlantic Ocean

Geographic context

Best-fitting function

(12) Amazon Slightly concave

(11) Zaire Slightly concave

System Longitudinal profile

Covault et al.

Canyon-and-channel longitudinal profiles 1985; Adams and Schlager, 2000). For the visual analysis, profiles were grouped into convex-upward, slightly concave-upward, and very concave-upward categories (Fig. 5). Hereafter, convex and concave will be used for simplification. For the curve-fitting analysis, we attempted to fit profiles to three simple functions, exponential, linear, and logarithmic, and we used coefficients of determination (r 2 values) from leastsquares regression to determine which function best describes a profile (cf. Shepherd, 1985; Adams and Schlager, 2000) (Fig. 2). Exponential functions are of the general form: y = ce bx,

(1)

where y is water depth, x is the down-system distance, c and b are constants, and e is the base of the natural logarithm. Exponential functions indicate that profiles change basinward at an increasing rate. Linear functions are of the general form: y = mx + b,

(2)

where m is the slope of the line of best fit, and b is the intersection of the line with the y axis. Logarithmic functions are of the general form: y = c ln x + b,

(3)

where c and b are constants, and ln is the natural logarithm. Logarithmic functions indicate that profiles change quickly and then level out basinward. They are the inverse of exponential functions. Simple visual analysis did not always correspond with more objective curve fitting. For example, visual inspection of the Nigeria X (2), San Antonio (4), Kushiro (5), and Aoga (6) profiles indicates that they are convex (Fig. 5); however, they are objectively best fit, according to least-squares regression, to linear functions, and, therefore, would be classified as linear profiles by the curve-fitting method of this study (Fig. 2).

Convex Profiles Six canyon-and-channel systems have convex profiles, which are relatively flat in their proximal reaches but are steeper in their distal reaches (East Breaks [1], Nigeria X [2], Barbados A [3], San Antonio [4], Kushiro [5], and Aoga [6]) (Fig. 6). Two of these profiles, East Breaks (1) and Barbados A (3), are best fit, according to least-squares regression, to exponential functions. The East Breaks (1) system is commonly referred to as the Brazos-Trinity system (e.g., Mallarino et al., 2006). Four profiles, Nigeria X (2), San Antonio (4), Kushiro (5), and Aoga (6), are best fit to linear functions (Fig. 2). These systems are from passive-margin slopes subjected to gravity-driven tectonic deformation that produces diapirism, growth faults, folds, and toe thrusts (e.g., the intraslope basin province of the western Gulf of Mexico for East Breaks [1] and offshore the Niger Delta continental margin for Nigeria X [2]; Damuth, 1994; Rowan et al., 2004) and tectonically active convergent margins (Barbados A [3], San Antonio [4], Kushiro [5], and Aoga [6]) (Table 1 and Fig. 6). All six of these systems were subjected to synsedimentary tectonic deformation and received relatively small volumes of sediment over the last glacial cycle relative to large submarine fan systems that were provided voluminous sediment in open ocean basins (discussed below in the Slightly Concave Profiles section) (cf. fluvial sediment-load measurements in Table 1). The development of depositional architecture on passive margins deformed by gravity-driven processes (East Breaks [1] and Nigeria X [2]) corresponds with subtle gradient changes across their diapiric and growth-faulted slopes (Pirmez et al., 2000) (Fig. 6C). Relatively fine-grained, shelf-edge, delta-fed sediment was transported through leveed channels of the small East Breaks (1) and Nigeria X (2) slope channels to pockets of intraslope accommodation, where ponded turbidite systems developed (Beaubouef and Friedmann, 2000; Booth et al., 2000; Pirmez et al., 2000). In contrast, the San Antonio (4) system offshore Chile and the Kushiro (5) and Aoga (6) systems offshore Hok-

kaido and Honshu, respectively, are relatively large canyon–to–erosional-channel systems that deposited relatively coarse-grained sediment in accretionary wedge-top and forearc basins, and transported sediment across the steep front of accretionary wedges that extend seaward into trenches (Klaus and Taylor, 1991; Laursen and Normark, 2002; Noda et al., 2008) (Table 1 and Fig. 6D). The Barbados A (3) system offshore Venezuela transported relatively coarse-grained sediment across the Barbados Ridge Complex (Huyghe et al., 2004) (Table 1). The system originates on a tectonically quiescent segment of the continental slope, where the gradient is flatter, which facilitated the development of levee and overbank relief. Across more distal reaches, however, the system lacks levee relief and is incised into the steeper, actively uplifting front of the Barbados Ridge Complex (Huyghe et al., 2004). Slightly Concave Profiles Seven canyon-and-channel systems have slightly concave profiles (Astoria [7], Rhone [8], Hudson [9], Mississippi [10], Zaire [11], Amazon [12], and Monterey [20]) (Fig. 7). All seven are best fit to linear functions (Fig. 2). Five of these systems (Rhone [8], Mississippi [10], Zaire [11], Amazon [12], and Monterey [20]) are associated with some of the largest deep-sea fans in the world (Barnes and Normark, 1985) (Table 1 and Fig. 7). These generally mud-rich systems include enormous canyons that transition to channels with well-developed levee and overbank relief and terminate as depositional lobes in some cases greater than one thousand kilometers down-system (Fig. 7). The Monterey (20) system is exceptional in that it developed across the California transform margin, and, similar to many of the sand-rich systems of the margin, was active during the Holocene marine transgression and highstand (Paull et al., 2005; Fildani et al., 2006; Piper and Normark, 2009). One slightly concave system, Astoria (7), developed across the Cascadia tectonically active convergent margin offshore western North America (Table 1).

Canyon-and-Channel Longitudinal Profiles Longitudinal profiles are presented below according to groups determined from visual inspection (Fig. 5). Information pertaining to best-fitting functions of profiles is also provided, as well as continental-margin and depositional-architecture context. Table 1 provides more information pertaining to the characteristics of canyon-and-channel systems and continental margins.

Figure 2 (on following five pages). Plots of all canyon-and-channel longitudinal profiles. Left: Lengths of entire profiles. Arrows indicate bases of continental slopes or ends of confined reaches of canyon-and-channel systems used for normalization. Bases of slopes were not measured at arbitrary or unit lengths, but were defined at the points where profiles reached gradients <0.25° and curvatures (i.e., down-system change in gradient) between –10–7 and 10–7. Notice distal reaches of only a few profiles were cut off by this normalization procedure (i.e., Mississippi [10], Zaire [11], and Amazon [12]). Right: Normalized profiles across continental slopes with best-fitting curves and functions according to least-squares regression (red lines).

Geosphere, April 2011

317

DOWN-SYSTEM LENGTH (m)

DOWN-SYSTEM LENGTH (m)

0

0

10000 20000 30000 40000 50000 60000 70000

200

(1) East Breaks

400 600 800

Intraslope end of system

1000 1200 1400 Normalized length

WATER DEPTH (m)

600 800 1000 1200

y = 436.29e 2E - 05x R² = 0.9832 Exponential

1400 1600

DOWN-SYSTEM LENGTH (m) 0

40000

60000

80000

(2) Nigeria X

400 600 800 1000 1200

Intraslope end of system

1400 1600 1800

WATER DEPTH (m)

20000

Normalized length

2000

20000

40000

60000

80000

0 200

(2) Nigeria X

400 600 800 1000 1200 1400 1600

y = 0.0185x + 386.95 R² = 0.9949 Linear

1800 2000

(see right column)

DOWN-SYSTEM LENGTH (m)

DOWN-SYSTEM LENGTH (m)

0

0

50000 100000 150000 200000 250000 300000

0 500

(3) Barbados A

1000 1500 2000

Base of slope and end of system

2500 3000 3500

Normalized length (see right column)

4000

50000 100000 150000 200000 250000 300000

0 500

(3) Barbados A

1000 1500 2000 2500

y = 1478.4e 3E - 06x R² = 0.9874 Exponential

3000 3500 4000

DOWN-SYSTEM LENGTH (m)

DOWN-SYSTEM LENGTH (m)

0

0

20000

40000

60000

80000 100000 120000

0

(4) San Antonio

2000 3000

Base of slope and end of system

4000 Normalized length (see right column)

6000

20000

40000

60000

80000 100000 120000

0 1000

(4) San Antonio

2000 3000 4000 5000 6000

Figure 2.

318

(1) East Breaks

400

DOWN-SYSTEM LENGTH (m) 200

5000

200

0 0

1000

10000 20000 30000 40000 50000 60000 70000

0

1800

(see right column)

WATER DEPTH (m)

WATER DEPTH (m)

1600

WATER DEPTH (m)

WATER DEPTH (m)

0

WATER DEPTH (m)

WATER DEPTH (m)

Covault et al.

Geosphere, April 2011

y = 0.0417x + 316.96 R² = 0.9753 Linear

Canyon-and-channel longitudinal profiles

DOWN-SYSTEM LENGTH (m) 50000

100000

150000

200000

0 1000

(5) Kushiro

2000 3000 4000

DOWN-SYSTEM LENGTH (m) 0

250000

Base of slope and end of system

5000 6000

WATER DEPTH (m)

WATER DEPTH (m)

0

Normalized length

7000

1000

150000

200000

(6) Aoga

2000 3000 4000

Base of slope and end of system

5000 6000

4000 5000

1000

(7) Astoria

1500 Base of slope

2500 Normalized length

3000

(see right column)

5000

50000

100000

150000

200000

250000

0 500

(7) Astoria

1000 1500 2000 2500

y = 0.013x + 463.34 R² = 0.9593 Linear

3000

DOWN-SYSTEM LENGTH (m) 50000

100000

150000

0

(8) Rhone

1000 Base of slope and end of system

1500 2000 Normalized length (see right column)

DOWN-SYSTEM LENGTH (m) 0

200000

WATER DEPTH (m)

WATER DEPTH (m)

y = 0.0273x + 711.33 R² = 0.9818 Linear

6000

3500

0

2500

250000

4000

0

3500

500

200000

DOWN-SYSTEM LENGTH (m) WATER DEPTH (m)

WATER DEPTH (m)

0

2000

150000

3000

50000 100000 150000 200000 250000 300000

1000

100000

(6) Aoga

2000

DOWN-SYSTEM LENGTH (m)

500

50000

0

7000

(see right column)

0

y = 0.03x - 34.786 R² = 0.9833 Linear

6000

0

Normalized length

7000

250000

3000

250000

0 1000

200000

DOWN-SYSTEM LENGTH (m) WATER DEPTH (m)

WATER DEPTH (m)

100000

150000

7000

(see right column)

50000

100000

(5) Kushiro

2000

DOWN-SYSTEM LENGTH (m) 0

50000

0

50000

100000

150000

200000

0 500

(8) Rhone

1000 1500 2000 2500

y = 0.0106x + 601.82 R² = 0.9259 Linear

3000

Figure 2 (continued).

Geosphere, April 2011

319

DOWN-SYSTEM LENGTH (m)

DOWN-SYSTEM LENGTH (m)

0

0

100000

200000

300000

0 500

(9) Hudson

1000 1500 2000 2500

Base of slope

3000 3500 Normalized length

4000

WATER DEPTH (m)

WATER DEPTH (m)

Covault et al.

1500 2000 2500 3000 3500

100000

500

200000

300000

400000

(10) Mississippi

1000 1500 Base of slope

2000 2500

y = 0.0134x + 785.12 R² = 0.9238 Linear

4000

DOWN-SYSTEM LENGTH (m)

Normalized length

0

WATER DEPTH (m)

WATER DEPTH (m)

0

300000

(9) Hudson

1000

DOWN-SYSTEM LENGTH (m) 0

200000

500

4500

(see right column)

100000

0

(see right column)

100000

200000

300000

0 500

(10) Mississippi

1000 1500 2000 y = 0.0086x + 317.93 R² = 0.9925 Linear

2500 3000

3000

DOWN-SYSTEM LENGTH (m)

DOWN-SYSTEM LENGTH (m) 200000

600000

1000000

(11) Zaire

1000

Base of slope

2000 3000 Normalized length (see right column)

4000

End of system

5000

0

WATER DEPTH (m)

WATER DEPTH (m)

0 0

1000 1500 2000

3000

DOWN-SYSTEM LENGTH (m)

400000

600000

800000 1000000

(12) Amazon Base of slope

2500 3000

End of system

3500 Normalized length (see right column)

100000

200000

300000

400000

0 500 1000

(12) Amazon

1500 2000 2500 3000 3500 4000

Figure 2 (continued).

320

y = 0.0072x + 353.51 R² = 0.9718 Linear

2500

0

200000

1500

4000

(11) Zaire

500

0

500

2000

300000

3500

0

1000

200000

DOWN-SYSTEM LENGTH (m) WATER DEPTH (m)

WATER DEPTH (m)

6000

100000

0

Geosphere, April 2011

y = 0.0081x + 600.98 R² = 0.9773 Linear

500000

DOWN-SYSTEM LENGTH (m)

DOWN-SYSTEM LENGTH (m)

0

0

300000

400000

0 500

(13) Laurentian

1000 1500 2000 2500 3000

Base of slope

3500 4000 4500

Normalized length

400000

(13) Laurentian y = 888ln(x) - 6658.9 R² = 0.9229 Logarithmic

DOWN-SYSTEM LENGTH (m) 0

10000

20000

200

30000

40000

50000

60000

(14) La Jolla

400 600 End of system

800 1000 Normalized length

10000

20000

30000

40000

50000

60000

0 200

(14) La Jolla

400 y = 284.55ln(x) - 2084.9 R² = 0.9643 Logarithmic

600 800 1000 1200

(see right column)

DOWN-SYSTEM LENGTH (m) 5000

10000

15000

20000

25000

DOWN-SYSTEM LENGTH (m)

0 100

(15) Carlsbad

200 300 400 500

End of system

600 700

Normalized length

0

30000

WATER DEPTH (m)

0

5000

10000

15000

20000

25000

0 100

(15) Carlsbad

200 300

y = 216.98ln(x) - 1500.2 R² = 0.9901 Logarithmic

400 500 600 700 800

(see right column)

800

DOWN-SYSTEM LENGTH (m)

DOWN-SYSTEM LENGTH (m)

0

0

20000

40000

60000

0

(16) Oceanside

400 600 End of system

800 1000 Normalized length

WATER DEPTH (m)

WATER DEPTH (m)

300000

DOWN-SYSTEM LENGTH (m)

1200

WATER DEPTH (m)

200000

0 0

200

100000

0 500 1000 1500 2000 2500 3000 3500 4000 4500 5000

(see right column)

5000

WATER DEPTH (m)

200000

WATER DEPTH (m)

100000

WATER DEPTH (m)

WATER DEPTH (m)

Canyon-and-channel longitudinal profiles

20000

40000

60000

0 200 400 600

(16) Oceanside y = 270.8ln(x) - 1914.6 R² = 0.9844 Logarithmic

800 1000

(see right column)

1200

1200

Figure 2 (continued).

Geosphere, April 2011

321

DOWN-SYSTEM LENGTH (m)

DOWN-SYSTEM LENGTH (m)

0

0

60000

80000 100000 120000

200

(17) Newport

400 600 Base of slope

800

End of system

1000 Normalized length

WATER DEPTH (m)

40000

0

20000

40000

60000

80000

100000

0 200

(17) Newport

400 y = 206.05ln(x) - 1419.2 R² = 0.9897 Logarithmic

600 800 1000

(see right column)

1200

WATER DEPTH (m)

20000

1200

DOWN-SYSTEM LENGTH (m)

DOWN-SYSTEM LENGTH (m)

0

0

50000

100000

150000

0 500

(18) Var

1000 1500 Base of slope End of system

2000 2500

Normalized length

WATER DEPTH (m)

WATER DEPTH (m)

Covault et al.

20000

40000

60000

80000

100000

0 500

(18) Var

1000 y = 515.19ln(x) - 3454.6 R² = 0.9743 Logarithmic

1500 2000 2500

(see right column)

3000

DOWN-SYSTEM LENGTH (m)

DOWN-SYSTEM LENGTH (m)

0

0

20000

40000

60000

80000

100000

0 500

(19) Ascension

1000 1500 2000

End of system

2500 3000

Normalized length

WATER DEPTH (m)

WATER DEPTH (m)

3000

20000

40000

60000

80000

100000

0 500

(19) Ascension

1000 1500

y = 750.06ln(x) - 5543.3 R² = 0.9847 Logarithmic

2000 2500 3000

(see right column)

DOWN-SYSTEM LENGTH (m)

DOWN-SYSTEM LENGTH (m)

0

0

50000

100000

150000

0 500

(20) Monterey

1000 1500 2000 2500

Base of slope

3000

WATER DEPTH (m)

WATER DEPTH (m)

3500

500 1000

2000 2500 3000 3500

(see right column)

4000

Figure 2 (continued).

322

100000

(20) Monterey

1500

Normalized length

3500

50000

0

Geosphere, April 2011

y = 0.0248x + 365.41 R² = 0.9452 Linear

150000

Canyon-and-channel longitudinal profiles

DOWN-SYSTEM LENGTH (m) 0

400000

800000

1200000

0 15

16

1000

14 17 1

WATER DEPTH (m)

2 8

2000

3000

10 18

(1) East Breaks

(8) Rhone

(15) Carlsbad

(2) Nigeria X

(9) Hudson

(16) Oceanside

(3) Barbados A

(10) Mississippi

(17) Newport

(4) SanAntonio

(11) Zaire

(18) Var

(5) Kushiro

(12) Amazon

(19) Ascension

(6) Aoga

(13) Laurentian

(20) Monterey

(7) Astoria

(14) La Jolla

7

19

3 20

4000

9 11 13

12

5000 4 5

6000 6

7000

Figure 3. Canyon-and-channel longitudinal profiles.

Very Concave Profiles Seven canyon-and-channel systems have very concave profiles (Laurentian [13], La Jolla [14], Carlsbad [15], Oceanside [16], Newport [17], Var [18], and Ascension [19]), which are relatively steep in their proximal reaches but are flat in their distal reaches (Fig. 8). All seven profiles are best fit to logarithmic functions (Fig. 2). Six of these systems are rich in sand and developed across the California and French Mediterranean margins. These margins are characterized by relatively steep slopes outboard of narrow shelves and nearby hinterlands from which relatively coarse-grained sediment is shed (La Jolla [14], Carlsbad [15], Oceanside [16], Newport [17], Var [18], and Ascension [19]) (Table 1 and Fig. 8). Five of these systems, La Jolla (14), Carlsbad (15), Oceanside (16), Newport (17), and Var (18), have canyons that transition to channels with modest levee and overbank relief and terminate as depositional lobes. The Ascension (19) system offshore central California also exhibits a welldeveloped canyon-and-channel system that does not terminate as depositional lobes; rather, it is a significant tributary to the Monterey (20) system (Normark et al., 1985c; Nagel et al., 1986; Greene et al., 2002; Fildani and Normark, 2004). The six aforementioned systems were subjected to synsedimentary tectonic deformation; however, the Laurentian (13) system developed across the Atlantic passive margin

and, similar to the Astoria (7) system discussed above, was fed relatively large volumes of coarse-grained sediment during subglacial transitions (Skene and Piper, 2006; Piper et al., 2007; Piper and Normark, 2009) (Table 1). The Laurentian (13) conduit is remarkable for its straightness, 25-km width, residual buttes, flat erosional floor, and spillover channels (Piper and Normark, 2009). Development of Longitudinal Profiles and Their Relationships to Continental Margin Types and Depositional Styles Because we examined canyon-and-channel systems on the modern seafloor, they reflect processes and forcings that operated since the last glacial cycle (e.g., <100 ka; Lambeck and Chappell, 2001). However, effects of the more distant past—for example, millions of years of sedimentary-basin filling or uplift of an accretionary wedge—also had a profound influence on some of these recently developed sedimentdelivery systems by establishing the seafloor template and morphology across which canyons and channels developed. The groups of normalized canyon-and-channel longitudinal profiles of this study generally reflect varying degrees of seafloor uplift and deformation, construction of depositional relief, and degradation of the seafloor by erosion associated with sediment gravity flows and other mass movements (Fig. 9). These influences are explained below in the

Geosphere, April 2011

context of their relative contributions to longitudinal-profile shapes in the convex, slightly concave, and very concave profile groups. There are only three canyon-and-channel systems whose longitudinal-profile shapes do not correspond well with continental-margin types or depositional architectures characteristic of other systems in their common profile group. These are the Astoria (7), Laurentian (13), and Monterey (20) systems, which are discussed below in the Exceptional Canyon-and-Channel Systems section. Convex profiles appear to have developed as a result of the dominance of seafloor uplift and deformation (e.g., East Breaks [1], Nigeria X [2]; Barbados A [3], San Antonio [4], Kushiro [5], and Aoga [6]) (Fig. 9). Such profiles developed in passive margins affected by gravity-driven tectonics and tectonically active convergent margins. Even though the signature of seafloor uplift and deformation is apparent in the convex shape of these profiles, other factors, such as erosion by sediment gravity flows, could have impacted seafloor morphology. Contraction above detachment surfaces in both types of belts can result in a broad zone of uplift and deformation, which is manifested in the evolving wedge shape of fold-and-thrust belts (Dahlen et al., 1984; Rowan et al., 2004). This evolving wedge shape maintains an approximately convex regional profile, which is also reflected by canyon-and-channel longitudinal profiles (Fig. 9). In unstable progradational, or supplydominated (Carvajal et al., 2009), passivemargin slope settings, gravitational instabilities can facilitate gravity-driven diapirism, growth faulting, and fold-and-thrust–related uplift and deformation (e.g., the western Gulf of Mexico for East Breaks [1] and offshore west Africa for Nigeria X [2]; Hedberg, 1970; Winker and Edwards, 1983; Rowan et al., 2004). In convergent margins, tectonic processes including basin-localized subsidence, fault-supported inner and outer margin uplift (e.g., Melnick et al., 2006; Collot et al., 2008), and construction of a frontal prism of accreted sediment (Dahlen et al., 1984; Rowan et al., 2004) can steepen the lower slope and work in concert with deep-sea canyon–and–channel-related sedimentary processes to produce the characteristic convex expression of longitudinal profiles (Ranero et al., 2006; von Huene et al., 2009). In particular, time-transgressive landward migration of the trench (e.g., Soh and Tokuyama, 2002; Noda et al., 2008), consequent margin steepening, and truncation of the submerged forearc caused by frontal subduction erosion can be fundamentally important to longitudinal profile development in convergent margins (Ranero et al., 2006; von Huene et al., 2009).

323

Covault et al.

A

NORMALIZED DOWN-SYSTEM LENGTH 0.0

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

0.9

1.0

0.0

0.2

NORMALIZED DEPTH

(1) East Breaks (2) Nigeria X (3) Barbados A (4) SanAntonio (5) Kushiro (6) Aoga

1

0.1

3 6

0.3

8

0.4

4

5

(8) Rhone (11) Zaire (12) Amazon (14) La Jolla (15) Carlsbad (16) Oceanside (17) Newport (19) Ascension

Relatively convex profiles

0.5

12 11

0.6

17

15 19

0.7

14

2

0.8 0.9

Relatively concave profiles

16

1.0

NORMALIZED DEPTH

0.0

NORMALIZED DOWN-SYSTEM LENGTH 0.0 0.2 0.4 0.6 0.8 1.0 (8) Rhone (11) Zaire (12) Amazon (14) La Jolla (15) Carlsbad (16) Oceanside (17) Newport (19) Ascension

0.2

8

0.4

12 0.6 0.8

17 15

11

14

19

0.0

NORMALIZED DEPTH

B

0.2 0.4

NORMALIZED DOWN-SYSTEM LENGTH 0.0 0.2 0.4 0.6 0.8 1.0 (8) Rhone (11) Zaire (12) Amazon (14) La Jolla (15) Carlsbad (16) Oceanside (17) Newport (19) Ascension

0.6 0.8

16 1.0

1.0

Figure 4. Canyon-and-channel longitudinal-profile normalization on the basis of profile length from canyon head to the end of the confined portion of the system (e.g., at the channel-to-lobe transition zone). (A) Only 14 profiles include the entire confined segments of their canyon-and-channel systems. (B) Closer inspection of relatively concave profiles shows that some profiles are distinctively more convex in their proximal reaches (dashed black box; <0.4 of their total down-system length) relative to other profiles.

Slightly concave profiles are commonly associated with mature passive continental margins not subjected to appreciable tectonic uplift, but gradual subsidence as a result of thermal cooling of the lithosphere (e.g., Rhone [8], Hudson [9], Mississippi [10], Zaire [11], and Amazon [12]) (Fig. 9). However, they can have pronounced relief as a result of preexisting depositional architecture. This preexisting depositional architecture is characteristic of constructional,

324

progradational, or supply-dominated margins, which have thick sedimentary prisms composed of well-developed clinothem and fan sequences (Hedberg, 1970; Ross et al., 1994; Carvajal et al., 2009; Gerber et al., 2009; Ryan et al., 2009a). Such preexisting depositional architecture establishes a relatively convex seafloor template across which canyons and channels extend. This seafloor template facilitates the development of a distinctively less concave

Geosphere, April 2011

proximal reach of mature passive-margin longitudinal profiles relative to profiles from the immature, underfilled margin offshore southern California characterized by steeper slopes and narrower shelves (Fig. 4B). Thus, progradation of these sediment supply–dominated passive margins associated with slope clinothem and fan accretion favors the development of a relatively convex seafloor across which channels subsequently incise and create a slightly concave profile shape (cf. Carvajal et al., 2009; Gerber et al., 2009; Ryan et al., 2009a). Very concave profiles predominantly developed on immature, or erosional (Ryan et al., 2009a), continental margins, some of which are dominated by strike-slip deformation (e.g., La Jolla [14], Carlsbad [15], Oceanside [16], Newport [17], Var [18], and Ascension [19]) (Fig. 9). In such settings, deformation contributes to steep slopes outboard of narrow shelves (cf. erosional margins of Ryan et al., 2009a). The narrow shelves offshore California and the French Riviera and nearby hinterlands suggest that relatively coarse-grained sediment is continuously fed to canyon-and-channel systems during both high and low stands of sea level (Savoye et al., 1993; Covault et al., 2007). Gerber et al. (2009) demonstrated the influence of erosion by sediment gravity flows across steep slopes of the Catalan margin in the Mediterranean Sea on the development longitudinal-profile concavity. Because the degree of degradation is related to the cumulative shear stress imposed by a number of sediment gravity flows on the seafloor (Middleton and Southard, 1984; Leeder, 1999; Boggs, 2001), relatively coarse grain sizes, more continuous erosion by sediment gravity flows, and steep slopes inherent to these immature margins promoted more significant seafloor degradation than in other settings (cf. Gerber et al., 2009). Exceptional Canyon-and-Channel Systems There are a couple of exceptional canyon-andchannel systems documented in this study that exhibit more concave profiles than one might predict on the basis of their continental-margin setting: the Astoria (7) and Laurentian (13) systems (Figs. 7, 8, and 10). Both are relatively high-latitude systems and, as a result, were particularly sensitive to climatic variability associated with the latest Pleistocene-to-Holocene glacial-to-interglacial transition (Table 1 and Fig. 10). The slightly concave, rather than the expected convex, profile of the Astoria (7) system might have resulted from the margin receiving pulses of coarse-grained sediment and water from periodic catastrophic floods of the Columbia River since the Last Glacial Maximum (Nelson et al., 1970; Piper and Normark, 2009).

Canyon-and-channel longitudinal profiles

NORMALIZED ACROSS-SLOPE LENGTH

A 0.0

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

0.9

1.0

0.0

NORMALIZED DEPTH

0.2

3 6

0.3 0.4 0.5 0.6

(1) East Breaks (2) Nigeria X (3) Barbados A (4) SanAntonio (5) Kushiro (6) Aoga (7) Astoria

1

0.1

9

8 11

13

4 7

15

0.7

10 2

14 WATER DEPTH (km)

0.8 0.9 1.0

0

0

(15) Carlsbad (16) Oceanside (17) Newport (18) Var (19) Ascension (20) Monterey

5

12

17

(8) Rhone (9) Hudson (10) Mississippi (11) Zaire (12) Amazon (13) Laurentian (14) La Jolla

LENGTH (km) 400 800

19 16 20

2 4

18

6

NORMALIZED ACROSS-SLOPE LENGTH

B 0.0

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

0.9

1.0

0.0 0.1

Convex profiles Slightly concave profiles

NORMALIZED DEPTH

0.2

Very concave profiles

0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0

Figure 5. (A) All canyon-and-channel longitudinal profiles normalized across their continental slopes. Bases of slopes were not measured at arbitrary or unit lengths; they were defined at the points where profiles reached gradients <0.25° and curvatures between –10–7 and 10–7. Inset: Full-length profiles from Figure 3. (B) Profile groups based on relative convexity or concavity from visual inspection.

Geosphere, April 2011

A similar situation exists for the Laurentian (13) system, which exhibits a very concave profile, even though it developed across the Atlantic passive margin (Table 1; Figs. 8 and 10B). For both high-latitude systems, numerous sandy and coarser-grained, thick mass movements initiated during glacial-to-interglacial transitions, which promoted seafloor erosion and the development of relatively concave longitudinal profiles (Skene and Piper, 2006; Piper et al., 2007; Piper and Normark, 2009). The Monterey (20) system is unique relative to other slightly concave profiles in that it is sand rich and developed across the California transform margin (Table 1 and Fig. 7). However, the Monterey (20) system feeds a large deepsea fan (Fildani and Normark, 2004). Voluminous preexisting fan deposits likely fostered a relatively convex proximal segment of the Monterey (20) profile, and turbidite deposition was recently focused in the proximal reaches of the canyon (Paull et al., 2005) (Figs. 2 and 7). The distal reaches of the profile are more steeply concave, which is common in strike-slip settings. Gerber et al. (2009) related such a convexto-concave longitudinal-profile shape across the Ebro margin in the Mediterranean Sea to sediment–gravity-flow deposition in a prograding canyon. Similarly, the Monterey (20) profile shows that the seafloor might be raised as a result of very recent localized deposition (i.e., during the Holocene, <10 ka; Lambeck and Chappell, 2001; Paull et al., 2005; Fildani et al., 2006). These observations suggest that canyons and channels are not merely conduits for sediment transport, but can be sites of deposition that might be detectable from profile analysis. Tectonic circumstances have also been interpreted to govern Monterey (20) profile morphology (Greene et al., 2002). Monterey Bay of the central California continental margin includes two contrasting physiographic and tectonic provinces separated by the Palo Colorado–San Gregorio fault zone: (1) the eastern Salinian Block comprises metamorphic and granitic plutonic rocks; and (2) the western San Simeon Block comprises Franciscan volcanic, metamorphic, and sedimentary rocks (Mullins and Nagel, 1981; Greene et al., 2002). The Palo Colorado–San Gregorio fault zone crosses the axis of Monterey (20) Canyon at ~2000-m water depth, which approximately coincides with the inflection point between relatively convex and concave segments of the canyon-and-channel system (Paull et al., 2005) (Fig. 2). Greene et al. (2002) noted different Monterey (20) Canyon morphologies across the fault zone: the eastern Salinian Block displays steeper, V-shaped canyon cross-sectional morphology; the western San Simeon Block displays broader, U-shaped

325

Covault et al.

DOWN-SYSTEM LENGTH (m)

A 0

0

400000

2000 0

1200000

C

United States Gulf Coast 30°N

(1) East Breaks Passive-margin slopes with (2) Nigeria X gravity-driven diapirism and fold-and-thrust deformation

2

Mississippi Canyon East Breaks canyon-and-channel system

1 0

400000

800000

(3) Barbados A (4) SanAntonio (5) Kushiro (6) Aoga

1000

1200000

Tectonically active convergent margins

3000

3

25°N

00

6 5

4000

0m –2 –3000 m000 m

2000

–1

WATER DEPTH (m)

1000

800000

4

0

80

160 MILES

5000 0

6000

95°W

80

160 KILOMETERS

90°W

85°W

7000

NORMALIZED DOWN-SYSTEM LENGTH

B

0.0

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

0.9

D

Hokkaido, Japan

1.0

0.0

NORMALIZED DEPTH

0.1

1

0.2 0.3

3

0.4

5

0.5 0.6 0.7

Kushiro Canyon-and-channel system

(1) East Breaks (2) Nigeria X (3) Barbados A (4) SanAntonio (5) Kushiro (6) Aoga

0m 00 –1 00 m 0 –2 0m 00 –3 m 00 –40 0 m 00 m –5 m 0 00 000 –6 –7

42°N

4

Convex profiles

2

6

0.8 0

0.9 1.0

0

142°E

144°E

146°E

30 30

60 MILES 60 KILOMETERS

148°E

Figure 6. Convex longitudinal profiles. (A) Profiles before normalization. (B) Profiles normalized across their continental slopes. (C) Example bathymetry from the Gulf of Mexico passive margin dominated by diapirism and gravity-driven, fold-and-thrust-related tectonic deformation. Contour interval is –1000 m. East Breaks (1) system is yellow (Pirmez et al., 2000). Bathymetry from GeoMapApp, http://www.geomapapp.org (Ryan et al., 2009b). (D) Example bathymetry from the Tokachi-oki forearc basin and accretionary wedge offshore Hokkaido, Japan. Contour interval is –1000 m. Kushiro (5) system is yellow (Noda et al., 2008). Bathymetry from GeoMapApp, http://www.geomapapp.org (Ryan et al., 2009b).

morphology. Paull et al. (2005) also noted that bends in the upper Monterey (20) Canyon are oriented parallel to regional structures, presumably as a result of differential erosion along fault-generated weaknesses (Greene, 1990). Methods of Grouping Longitudinal Profiles We employed two methods in order to group longitudinal profiles: (1) visual inspection and categorization on the basis of relative convexity or concavity; and (2) more objective categorization on the basis of best-fitting mathematical functions (cf. Shepherd, 1985; Adams and

326

Schlager, 2000). Profiles that were grouped based on visual inspection generally correspond with continental-margin type and depositional architecture, with a few exceptions. More objective curve-fitting methods do an adequate job of differentiating the most convex and concave profiles; however, there is more ambiguity associated with differentiating between less convex and slightly concave profiles (Fig. 2). For example, visual inspection of the Nigeria X (2), San Antonio (4), Kushiro (5), and Aoga (6) profiles indicates that they are convex (Fig. 5); however, they are objectively best fit to linear functions and, therefore, are linear profiles along with the

Geosphere, April 2011

Rhone (8), Hudson (9), Mississippi (10), Zaire (11), and Amazon (12) profiles (Fig. 2). The Nigeria X (2), San Antonio (4), Kushiro (5), and Aoga (6) systems developed in the Niger Delta passive margin affected by gravity-driven tectonics and tectonically active convergent margins, and contributed to distinctively different depositional architectures relative to the Rhone (8), Hudson (9), Mississippi (10), Zaire (11), and Amazon (12) systems. More objective numerical methods, however, have been shown to adequately differentiate submarine geomorphic features. Adams and Schlager (2000) grouped 19 passive continental-

Canyon-and-channel longitudinal profiles

A –4000 m –30

00 m –2000 m –1000 m

Amazon Canyon-and-channel system

5°N

Brazil

0

100

0

55°W

50°W



200 KILOMETERS

45°W

DOWN-SYSTEM LENGTH (m)

B 0

0

400000

8

2000

800000 (8) Rhone (9) Hudson (10) Mississippi (11) Zaire (12) Amazon

1000

WATER DEPTH (m)

100

200 MILES

1200000

Mature passive margins

10

3000

7 9

20

4000

11 12

5000

(7) Astoria (20) Monterey

Exceptional systems

6000 7000

C

NORMALIZED DOWN-SYSTEM LENGTH

0.0 0.2

0.0

0.2

8

0.4

0.6

0.8

(8) Rhone (11) Zaire (12) Amazon

0.4 0.6

12 11 0.8 1.0

Slightly concave profiles

1.0

NORMALIZED DEPTH

NORMALIZED DEPTH

(Normalized from canyon head to end of channel)

(Normalized across slope) 0.0

0.0

0.2

0.4

0.2 0.4

10

0.6 0.8

9

12

0.6

0.8

1.0

(7) Astoria (8) Rhone (9) Hudson (10) Mississippi (11) Zaire (12) Amazon (20) Monterey

7

20 Slightly concave profiles

8 11

1.0

Figure 7. Slightly concave longitudinal profiles. (A) Example bathymetry from the Amazon deep-sea fan (GeoMapApp, http:// www.geomapapp.org; Ryan et al., 2009b). Contour interval is –1000 m. Amazon (12) system is yellow (Pirmez and Imran, 2003). (B) Profiles before normalization. (C) Left: Profiles normalized from canyon heads to ends of channels. Right: Profiles normalized across their continental slopes.

Geosphere, April 2011

327

Covault et al.

A -1000

m

-2000 m

44°N

France

Var Canyon-and-channel system

43°N

0

30

0

6°E

7°E

Sout

B

hern

30

60 MILES 60 KILOMETERS

9°E

8°E 0

Calif

ornia

10°E 10

0

10

20 MILES 20 KILOMETERS

-10

0m

-500

m

tem

d sba

sys

l

Car

O

ce an si de

sy st em

Newport system m

33°N 118°W

117°W

DOWN-SYSTEM LENGTH (m)

C WATER DEPTH (m)

Figure 8. Very concave longitudinal profiles. (A) Example bathymetry from the Mediterranean Sea (GeoMapApp, http://www.geomapapp.org; Ryan et al., 2009b). Contour interval is –1000 m. Var (18) system is yellow (Savoye et al., 1993). Notice proximity to Alpine sediment source area, narrow shelf, and steep slope. (B) Example bathymetry from offshore southern California (Gardner and Dartnell, 2002; Divins and Metzger, 2009). Contour interval is –100 m. Carlsbad (15), Oceanside (16), and Newport (17) systems are yellow (Graham and Bachman, 1983; Covault et al., 2007; Covault et al., 2010). (C) Profiles before normalization. (D) Left: Profiles normalized from canyon heads to ends of channels. Right: Profiles normalized across their continental slopes.

0

0

400000

15 16 17 16

1000 2000

18 19

3000 4000

D

800000

(14) La Jolla (15) Carlsbad (16) Oceanside (17) Newport (18) Var (19) Ascension

(13)Laurentian

Young margins with steep slopes and narrow shelves

Exceptional system

13

NORMALIZED DOWN-SYSTEM LENGTH

328

0.0

0.2

0.4

0.6

0.2 0.4 0.6

0.8

1.0

(14) La Jolla (15) Carlsbad (16) Oceanside (17) Newport

17

14 16

0.8 Very concave profiles

1.0

Geosphere, April 2011

15

NORMALIZED DEPTH

NORMALIZED DEPTH

(Normalized from canyon head to end of channel)

0.0

1200000

(Normalized across slope)

0.0

0.0

0.2

0.4

0.6

0.2 0.4 0.6

1.0

14 15 19

0.8 1.0

0.8

(13) Laurentian (14) La Jolla (15) Carlsbad (16) Oceanside (17) Newport (18) Var (19) Ascension

Very concave profiles

18 13

16 17

Canyon-and-channel longitudinal profiles

Dominant control Long-profile shape NORMALIZED ACROSS-SLOPE LENGTH

NORMALIZED DEPTH

Preexisting depositional relief

0.0 0.0

1.0

Margin profile and architecture Slightly concave Appreciable pre-existing depositional relief associated with mature margins

Extensive leveed channel

Slightly concave

Preexisting fan deposits

1.0

NORMALIZED ACROSS-SLOPE LENGTH

0.0 0.0 NORMALIZED DEPTH

Tectonic uplift and deformation

1.0

Convex Tectonic uplift and deformation in passive-margin gravity-driven and convergent margins

Turbidites in pockets of accommodation

Convex

1.0

NORMALIZED ACROSS-SLOPE LENGTH

0.0 0.0 NORMALIZED DEPTH

Erosional sedimentary processes

1.0

Very concave Enhanced sediment-gravity-flow erosion associated with young, underfilled margins with steep slopes and close proximity to source areas

Very concave

Narrow shelf

Nearby hinterland

Sediment gravity flow accelerating down the steep slope Coarse-grained fan

1.0

Figure 9. Summary chart of controls on longitudinal-profile shape: preexisting depositional relief, tectonic uplift and deformation, and erosional sedimentary processes. Although profiles reflect varying degrees of controlling factors, and exceptional profiles exist, profile groups are located adjacent to dominant controls and general continental-margin types with characteristic depositional architectures. Average profiles for each group are represented by bold black lines.

margin profiles on the basis of three functions: linear, exponential, and Gaussian distribution. Linear functions describe unstable margins that are interpreted by Adams and Schlager (2000) to rest at the angle of repose. Exponential functions represent the exponential decay of sediment transport capacity or competence with increasing distance across a margin (Adams and Schlager, 2000). The majority of margins fol-

low a Gaussian curve as a result of perturbing extrinsic processes at the shelf edge (e.g., wavereworking processes and shelf-edge instabilities; Adams and Schlager, 2000). Quantitative characteristics of margins are effective at predicting depositional architecture and sediment caliber (Adams and Schlager, 2000). Why were objective curve-fitting methods of entire-margin profiles more effective at creating meaningful

Geosphere, April 2011

and predictive groups relative to similar methods applied to canyon-and-channel longitudinal profiles of this study? Adams and Schlager (2000) only examined passive continental margins in order to avoid tectonic influences on primary depositional setting (Pratson and Haxby, 1996). If we were to exclude canyons and channels of tectonically active continental margins from our analysis, profiles would be

329

Covault et al.

A

0 0

200 200

400 MILES

400 KILOMETERS

50°N Astoria Canyon-and-channel system

Western North America

Columbia River drainage basin

–1000 m –2000 m

130°W

45°N

120°W

110°W

B

0 0

Eastern North America

200 200

400 MILES

400 KILOMETERS

50°N Glacia lly

e xc

a te va

d

–1000

l ne an

an oti lf c S he s

ch

Laurentian system

0m

00

70°W

CONCLUSIONS

m

–5

60°W

40°N

50°W

Figure 10. High-latitude continental margins conducive to the development of exceptionally concave canyon-and-channel longitudinal profiles. (A) Western North America topography and offshore Cascadia margin bathymetry from GeoMapApp, http://www.geomapapp.org (Ryan et al., 2009b). Contour interval is –1000 m. Columbia River drainage basin is outlined in white. Astoria (7) system is a dashed white line offshore (Nelson et al, 1970). (B) Scotian margin bathymetry offshore eastern North America from GeoMapApp, http://www.geomapapp.org (Ryan et al., 2009b). Contour interval is –1000 m. Notice the prominent glacially excavated channel leading to the Laurentian (13) system (dashed white line; Skene and Piper, 2006).

330

Geosphere, April 2011

neatly differentiated into two groups: (1) linear profiles across mature margins with large depositional fans (e.g., Rhone [8], Hudson [9], Mississippi [10], Zaire [11], and Amazon [12]); and (2) logarithmic profiles across settings dominated by erosional sedimentary processes (e.g., Laurentian [13] and Var [18]). Visual inspection of longitudinal profiles, therefore, is relatively useful in differentiating profiles into groups with some predictive capability of continental-margin type and depositional architecture; however, exceptional longitudinal profiles dominated by erosional sedimentary processes (e.g., Astoria [7] and Laurentian [13]) show the shortcomings of simple visual inspection, and highlight the need for more rigorous and objective numerical methods for differentiating profiles. Gerber et al. (2009) developed a morphodynamic model that predicts submarine canyon-and-channel longitudinal profiles affected by continental-margin progradation and the downslope evolution of turbidity currents. Such a model is an important step toward assessing the relative contributions of forcings to longitudinal-profile character, including preexisting depositional relief, margin evolution, and erosional sedimentary processes. Subsequent modeling efforts should focus on balancing those effects with tectonic uplift and deformation (cf. Fig. 9). Such models can be used in order to more precisely and meaningfully differentiate profiles and test some of the interpretations of this study, which were developed from more empirical observations and quantification of profiles.

Canyon-and-channel systems on the modern seafloor, and their characteristic longitudinal profiles, reflect processes and forcings that operated approximately since the last glacial cycle, for example, sediment–gravity-flow erosion (Fig. 9). However, effects of the more distant past can also have a profound influence on more recent sediment-delivery systems by establishing the seafloor template across which canyons and channels develop (Fig. 9). Longitudinal profile groups generally correspond with continental-margin types and depositional architectures, with a few exceptions. Furthermore, our assessment of methods of grouping profiles indicates that empirical observations of profiles are an effective means of constructing groups that have some predictive capability of continentalmargin type and depositional architecture. However, exceptional longitudinal profiles dominated by erosional sedimentary processes highlight the need for more rigorous and objective numerical methods for differentiating profiles and

Canyon-and-channel longitudinal profiles assessing controls on their development. Results of this study provide a new catalog of the breadth and general controls of the shapes of submarine sediment-delivery systems, which can be related to the depositional architecture of different continental margins. ACKNOWLEDGMENTS

The authors would like to thank the Clastic Stratigraphy Research and Development Team at Chevron Energy Technology Company (ETC) and the larger family of Earth scientists at ETC. In particular, Julian Clark, Angela Hessler, Michael Pyrcz, Brad Ritts, Will Schweller, and Yongjun Yue shared ideas, comments, and stimulating discussions with the authors. We also thank George Hilley and Steve Hubbard for invaluable discussions and insight, and Bill Normark for his mentorship and inspiration. We also thank Ivano Aiello, John Damuth, Atsushi Noda, David Piper, and David Scholl for constructive reviews. REFERENCES CITED Adams, E.W., and Schlager, W., 2000, Basic types of submarine slope curvature: Journal of Sedimentary Research, v. 70, p. 814–828, doi: 10.1306/2DC4093A -0E47-11D7-8643000102C1865D. Babonneau, N., Savoye, B., Cremer, M., and Klein, B., 2002, Morphology and architecture of the present canyon and channel system of the Zaire deep-sea fan: Marine and Petroleum Geology, v. 19, p. 445–467, doi: 10.1016/ S0264-8172(02)00009-0. Barnes, N.E., and Normark, W.R., 1985, Diagnostic parameters for comparing modern submarine fans and ancient turbidite systems, in Bouma, A.H., Normark, W.R., and Barnes, N.E., eds., Submarine Fans and Related Turbidite Systems: New York, Springer-Verlag, p. 13–14. Beaubouef, R.T., and Friedmann, S.J., 2000, High resolution seismic/sequence stratigraphic framework for the evolution of the Pleistocene intra slope basins, western Gulf of Mexico: Depositional models and reservoir analogs: Gulf Coast Section SEPM (Society of Sedimentary Geology) 20th Annual Research Conference Proceedings, p. 40–60. Boggs, S., 2001, Principles of Sedimentology and Stratigraphy: Third Edition: Upper Saddles River, New Jersey, Prentice Hall, 726 p. Booth, J.R., DuVernay, A.E., III, Pfeiffer, D.S., and Styzen, M.J., 2000, Sequence stratigraphic framework, depositional models, and stacking patterns of ponded and slope fan systems in the Auger Basin: Central Gulf of Mexico slope: Gulf Coast Section SEPM (Society of Sedimentary Geology) 20th Annual Research Conference Proceedings, p. 82–103. Bouma, A.H., Normark, W.R., and Barnes, N.E., 1985, Submarine Fans and Related Turbidite Systems: New York, Springer-Verlag, 351 p. Butman, B., Twichell, D.C., Rona, P.A., Tucholke, B.E., Middleton, T.J., and Robb, J.M., 2004, Sea floor topography and backscatter intensity of the Hudson Canyon region offshore of New York and New Jersey: U.S. Geological Survey Open-File Report 2004-1441, http://pubs.usgs.gov/of/2004/1441/. Carvajal, C., Steel, R., and Petter, A., 2009, Sediment supply: The main driver of shelf-margin growth: EarthScience Reviews, v. 96, p. 221–248, doi: 10.1016/ j.earscirev.2009.06.008. Collot, J.-Y., Agudelo, W., Ribodetti, A., and Marcaillou, B., 2008, Origin of a crustal splay fault and its relation to the seismogenic zone and underplating at the erosional north Ecuador–south Colombia oceanic margin: Journal of Geophysical Research, v. 113, p. B12102, doi: 10.1029/2008JB005691. Covault, J.A., Normark, W.R., Romans, B.W., and Graham, S.A., 2007, Highstand fans in the California borderland: The overlooked deep-water depositional systems: Geology, v. 35, p. 783–786, doi: 10.1130/G23800A.1.

Covault, J.A., Romans, B.W., Fildani, A., McGann, M., and Graham, S.A., 2010, Rapid climatic signal propagation from source to sink in a southern California sediment-routing system: The Journal of Geology, v. 118, p. 247–259, doi: 10.1086/651539. Dahlen, F.A., Suppe, J., and Davis, D., 1984, Mechanics of fold-and-thrust belts and accretionary wedges: Cohesive Coulomb theory: Journal of Geophysical Research, v. 89, p. 10087–10101, doi: 10.1029/JB089iB12p10087. Damuth, J.E., 1994, Neogene gravity tectonics and depositional processes on the deep Niger Delta continental margin: Marine and Petroleum Geology, v. 11, p. 320– 346, doi: 10.1016/0264-8172(94)90053-1. Dartnell, P., Normark, W.R., Driscoll, N.W., Babcock, J.M., Gardner, J.V., Kvitek, R.G., and Iampietro, P.J., 2007, Multibeam Bathymetry and Selected Perspective Views Offshore San Diego, California: U.S. Geological Survey Scientific Investigations Map 2959, http:// pubs.usgs.gov/sim/2007/2959/. Divins, D.L., and Metzger, D., 2009, NGDC Coastal Relief Model: http://www.ngdc.noaa.gov/mgg/coastal/ coastal.html. Estrada, F., Ercilla, G., and Alonso, B., 2005, Quantitative study of a Magdalena submarine channel (Caribbean Sea): Implications for sedimentary dynamics: Marine and Petroleum Geology, v. 22, p. 623–635, doi: 10.1016/j.marpetgeo.2005.01.004. Fildani, A., and Normark, W.R., 2004, Late Quaternary evolution of channel and lobe complexes of Monterey Fan: Marine Geology, v. 206, p. 199–213, doi: 10.1016/ j.margeo.2004.03.001. Fildani, A., Normark, W.R., Kostic, S., and Parker, G., 2006, Channel formation by flow stripping: Large-scale scour features along the Monterey East Channel: Sedimentology, v. 53, p. 1265–1287, doi: 10.1111/j.1365 -3091.2006.00812.x. Flood, R.D., and Damuth, J.E., 1987, Quantitative characteristics of sinuous distributary channels on the Amazon deep-sea fan: Geological Society of America Bulletin, v. 98, p. 728–738, doi: 10.1130/0016-7606 (1987)98<728:QCOSDC>2.0.CO;2. Flood, R.D., Manley, P.L., Kowsmann, R.O., Appi, C.J., and Pirmez, C., 1991, Seismic facies and late Quaternary growth of Amazon submarine fan, in Weimer, P., and Link, M.H., eds., Seismic Facies and Sedimentary Processes of Submarine Fans and Turbidite Systems: New York, Springer-Verlag, p. 415–433. Gardner, J.V., and Dartnell, P., 2002, Multibeam mapping of the Los Angeles, California margin: U.S. Geological Survey Open-File Report 02-162, http://geopubs.wr .usgs.gov/open-file/of02-162/. Gerber, T.P., Amblas, D., Wolinsky, M.A., Pratson, L.F., and Canals, M., 2009, A model for the long-profile shape of submarine canyons: Journal of Geophysical Research, v. 114, p. F03002, doi: 10.1029/2008JF001190. Gilbert, G.K., 1880, Report on the geology of the Henry Mountains: U.S. Geographical and Geological Survey of the Rocky Mountain Region, 170 p. Goff, J.A., 2001, Quantitative classification of canyon systems on continental slopes and a possible relationship to slope curvature: Geophysical Research Letters, v. 28, p. 4359–4362, doi: 10.1029/2001GL013300. Graham, S.A., and Bachman, S.B., 1983, Structural controls on submarine-fan geometry and internal architecture: Upper La Jolla Fan system, offshore southern California: American Association of Petroleum Geologists Bulletin, v. 67, p. 83–96. Greene, H.G., 1990, Regional tectonics and structural evolution of the Monterey Bay region, central California, in Garrison, R.E., Green, H.G., Hicks, K.R., Weber, G.E., and Wright, T.L., eds., Geology and Tectonics of the Central California Coastal Region, San Francisco to Monterey: Pacific Section, American Association of Petroleum Geologists Guidebook 67, p. 31–56. Greene, H.G., Maher, N.M., and Paull, C.K., 2002, Physiography of the Monterey Bay National Marine Sanctuary and implications about continental margin development: Marine Geology, v. 181, p. 55–82, doi: 10.1016/ S0025-3227(01)00261-4. Hedberg, H.D., 1970, Continental margins from viewpoint of the petroleum geologist: American Association of Petroleum Geologists Bulletin, v. 54, p. 3–43.

Geosphere, April 2011

Huyghe, P., Foata, M., Deville, Mascle, G., and Caramba Working Group, 2004, Channel profiles through the active thrust front of the southern Barbados prism: Geology, v. 32, p. 429–432. Inman, D.L., 2008, Highstand fans in the California borderland: Comment: Geology, Online Forum, e166. Inman, D.L., and Jenkins, S.A., 1999, Climate change and the episodicity of sediment flux of small California rivers: The Journal of Geology, v. 107, p. 251–270, doi: 10.1086/314346. Klaus, A., and Taylor, B., 1991, Submarine canyon development in the Izu-Bonin forearc: A SeaMARC II and seismic survey of Aoga Shima Canyon: Marine Geophysical Researches, v. 13, p. 131–152. Lambeck, K., and Chappell, J., 2001, Sea level change through the last glacial cycle: Science, v. 292, p. 679– 686, doi: 10.1126/science.1059549. Laursen, J., and Normark, W.R., 2002, Late Quaternary evolution of the San Antonio Submarine Canyon in the Central Chile forearc (~33°S): Marine Geology, v. 188, p. 365–390, doi: 10.1016/S0025-3227(02)00421-8. Leeder, M., 1999, Sedimentology and sedimentary basins: From turbulence to tectonics: Oxford, Blackwell Science, 592 p. Mallarino, G., Beaubouef, R.T., Droxler, A.W., Abreu, V., and Labeyrie, L., 2006, Sea level influence on the nature and timing of a minibasin sedimentary fill (northwestern slope of the Gulf of Mexico): American Association of Petroleum Geologists Bulletin, v. 90, p. 1089–1119, doi: 10.1306/02210605058. Melnick, D., Bookhagen, B., Echtler, H.P., and Strecker, M.R., 2006, Coastal deformation and great subduction earthquakes, Isla Santa María, Chile (37°S): Geological Society of America Bulletin, v. 118, p. 1463–1480, doi: 10.1130/B25865.1. Menard, H.W., 1955, Deep-sea channels, topography, and sedimentation: American Association of Petroleum Geologists Bulletin, v. 39, p. 236–255. Middleton, G.V., and Southard, J.B., 1984, Mechanics of sediment movement: Second edition: SEPM (Society of Sedimentary Geology), Short Course 3, 401 p. Milliman, J.D., and Syvitski, J.P.M., 1992, Geomorphic/ tectonic control of sediment discharge to the ocean: The importance of small mountainous rivers: The Journal of Geology, v. 100, p. 525–544, doi: 10.1086/629606. Milliman, J.D., Rutkowski, C., and Meybeck, M., 1995, River discharge to the sea: A global river index (GLORI): Land-Ocean Interactions in the Coastal Zone Reports and Studies 34, 125 p. Mitchell, N.C., 2005, Interpreting long-profiles of canyons in the USA Atlantic continental slope: Marine Geology, v. 214, p. 75–99, doi: 10.1016/j.margeo.2004.09.005. Mullins, H.T., and Nagel, D.K., 1981, Franciscan-type rocks off Monterey Bay, California: Implications for western boundary of the Salinian block: Geo-Marine Letters, v. 1, p. 135–139, doi: 10.1007/BF02463331. Mutti, E., and Normark, W.R., 1987, Comparing examples of modern and ancient turbidite systems: Problems and concepts, in Leggett, J.K., and Zuffa, G.G., eds., Marine Clastic Sedimentology: Concepts and Case Studies: London, Graham and Trotman, p. 1–38. Mutti, E., and Normark, W.R., 1991, An integrated approach to the study of turbidite systems, in Weimer, P., and Link, M.H., eds., Seismic Facies and Sedimentary Processes of Submarine Fans and Turbidite Systems: New York, Springer-Verlag, p. 75–106. Nagel, D.K., Mullins, H.T., and Greene, H.G., 1986, Ascension submarine canyon, California—Evolution of a multi-head canyon system along a strike-slip continental margin: Marine Geology, v. 73, p. 285–310, doi: 10.1016/0025-3227(86)90019-8. Nelson, C.H., Carlson, P.R., Byrne, J.V., and Alpha, T.R., 1970, Development of the Astoria Canyon-Fan physiography and comparison with similar systems: Marine Geology, v. 8, p. 259–291, doi: 10.1016/0025 -3227(70)90047-2. Noda, A., TuZino, T., Furukawa, R., Joshima, M., and Uchida, J., 2008, Physiographical and sedimentological characteristics of submarine canyons developed upon an active forearc slope: The Kushiro submarine canyon, northern Japan: Geological Society of America Bulletin, v. 120, p. 750–767.

331

Covault et al. Normark, W.R., 1970, Growth patterns of deep-sea fans: American Association of Petroleum Geologists Bulletin, v. 54, p. 2170–2195. Normark, W.R., 1985, Local morphologic controls and effects of basin geometry on flow processes in deep marine basins, in Zuffa, G.G., ed., Provenance of Arenites: North Atlantic Treaty Organization Advanced Science Institutes Series C-148, p. 47–63. Normark, W.R., Barnes, N.E., and Bouma, A.H., 1985a, Comments and new directions for deep-sea fan research, in Bouma, A.H., Normark, W.R., and Barnes, N.E., eds., Submarine Fans and Related Turbidite Systems: New York, Springer-Verlag, p. 341–343. Normark, W.R., Barnes, N.E., and Coumes, F., 1985b, Rhone Fan, Mediterranean, in Bouma, A.H., Normark, W.R., and Barnes, N.E., eds., Submarine Fans and Related Turbidite Systems: New York, SpringerVerlag, p. 151–156. Normark, W.R., Gutmacher, C.E., Chase, T.E., and Wilde, P., 1985c, Monterey Fan, Pacific Ocean, in Bouma, A.H., Normark, W.R., and Barnes, N.E., eds., Submarine Fans and Related Turbidite Systems: New York, Springer-Verlag, p. 79–86. Normark, W.R., and Piper, D.J.W., 1991, Initiation processes and flow evolution of turbidity currents: Implications for the depositional record: SEPM (Society of Sedimentary Geology) Special Publication, v. 46, p. 207–230. Normark, W.R., Posamentier, H., and Mutti, E., 1993, Turbidite systems: State of the art and future directions: Reviews of Geophysics, v. 31, p. 91–116. Paull, C.K., Mitts, P., Ussler, William, Keaten, R., and Greene, H.G., 2005, Trail of sand in upper Monterey Canyon: Offshore California: Geological Society of America Bulletin, v. 117, p. 1134–1145. Piper, D.J.W., and Normark, W.R., 2009, Processes that initiate turbidity currents and their influence on turbidites: A marine geology perspective: Journal of Sedimentary Research, v. 79, p. 347–362, doi: 10.2110/jsr.2009.046. Piper, D.J.W., Shaw, J., and Skene, K.I., 2007, Stratigraphic and sedimentological evidence for late Wisconsinan subglacial outburst floods to Laurentian Fan: Palaeogeography, Palaeoclimatology, Palaeoecology, v. 246, p. 101–119, doi: 10.1016/j.palaeo.2006.10.029. Pirmez, C., and Imran, J., 2003, Reconstruction of turbidity currents in Amazon Channel: Marine and Petroleum Geology, v. 20, p. 823–849, doi: 10.1016/ j.marpetgeo.2003.03.005. Pirmez, C., Beaubouef, R.T., Friedmann, S.J., and Mohrig, D.C., 2000, Equilibrium profile and baselevel in submarine channels: Examples from late Pleistocene systems and implications for the architecture of deepwater reservoirs: Gulf Coast Section SEPM (Society of Sedi-

332

mentary Geology) 20th Annual Research Conference Proceedings, p. 782–805. Playfair, J., 1802, Illustrations of the Huttonian Theory of the Earth: London, William Creech, 528 p. Pratson, L.F., and Haxby, W.F., 1996, What is the slope of the U.S. continental slope?: Geology, v. 24, p. 3–6, doi: 10.1130/0091-7613(1996)024<0003:WITSOT> 2.3.CO;2. Pratson, L.F., Ryan, W.B.F., Mountain, G.S., and Twichell, D.C., 1994, Submarine canyon initiation by downslopeeroding sediment flows: Evidence in late Cenozoic strata on the New Jersey continental slope: Geological Society of America Bulletin, v. 106, p. 395–412, doi: 10.1130/0016-7606(1994)106<0395:SCIBDE> 2.3.CO;2. Ranero, C.R., von Huene, R., Weinrebe, W., and Reichert, C., 2006, Tectonic processes along the Chile convergent margin, in Oncken, O., et al., eds., The Andes: Active Subduction Orogeny: Berlin, Springer-Verlag, p. 91–121. Reading, H.G., and Richards, M., 1994, Turbidite systems in deep-water basin margins classified by grain size and feeder system: American Association of Petroleum Geologists Bulletin, v. 78, p. 792–822. Ross, W.C., Halliwell, B.A., May, J.A., Watts, D.E., and Syvitski, J.P.M., 1994, Slope readjustment: A new model for the development of submarine fans and aprons: Geology, v. 22, p. 511–514, doi: 10.1130/ 0091-7613(1994)022<0511:SRANMF>2.3.CO;2. Rowan, M.G., Peel, F.J., and Vendeville, B.C., 2004, Gravitydriven fold belts on passive margins: American Association of Petroleum Geologist Memoir, v. 82, p. 157–182. Ryan, M.C., Helland-Hansen, W., Johannessen, E.P., and Steel, R.J., 2009a, Erosional versus accretionary shelf margins: The influence of margin type of deepwater sedimentation: An example from the Porcupine Basin, offshore western Ireland: Basin Research, doi: 10.1111/j.1365-2117.2009.00424.x. Ryan, W.B.F., Carbottee, S.M., Coplan, J.O., O’Hara, S., Melkonian, A., Arko, R., Weissel, R.A., Ferrini, V., Goodwillie, A., Nitsche, F., Bonczkowski, J., and Zemsky, R., 2009b, Global multi-resolution topography synthesis: Geochemistry Geophysics Geosystems, v. 10, p. Q03014, doi: 10.1029/2008GC002332. Savoye, B., Piper, D.J.W., and Droz, L., 1993, Plio-Pleistocene evolution of the Var deep-sea fan off the French Riviera: Marine and Petroleum Geology, v. 10, p. 550– 571, doi: 10.1016/0264-8172(93)90059-2. Schumm, S.A., Dumont, J.F., and Holbrook, J.M., 2000, Active tectonics and alluvial rivers: Cambridge, Cambridge University Press, 276 p. Shanmugam, G., and Moiola, R.J., 1988, Submarine fans: Characteristics, models, classification, and reservoir

Geosphere, April 2011

potential: Earth-Science Reviews, v. 24, p. 383–428, doi: 10.1016/0012-8252(88)90064-5. Shepard, F.P., 1948, Submarine Geology: New York, Harper and Brothers, 338 p. Shepard, F.P., 1981, Submarine canyons: Multiple causes and long-time persistence: American Association of Petroleum Geologist Bulletin, v. 65, p. 1062–1077. Shepherd, R.G., 1985, Analysis of river profiles: The Journal of Geology, v. 93, p. 377–384, doi: 10.1086/628959. Skene, K.I., and Piper, D.J.W., 2006, Late Cenozoic evolution of Laurentian Fan: Development of a glacially-fed submarine fan: Marine Geology, v. 227, p. 67–92, doi: 10.1016/j.margeo.2005.12.001. Snow, R.S., and Slingerland, R.L., 1987, Mathematical modeling of graded river profiles: The Journal of Geology, v. 95, p. 15–33, doi: 10.1086/629104. Soh, W., and Tokuyama, H., 2002, Rejuvenation of submarine canyon associated with ridge subduction, Tenryu Canyon, off Tokai, central Japan: Marine Geology, v. 187, p. 203–220, doi: 10.1016/S0025-3227 (02)00267-0. Sømme, T.O., Helland-Hansen, W., Martinsen, O.J., and Thurmond, J.B., 2009, Relationship between morphological and sedimentological parameters in source-tosink systems: A basis for predicting semi-quantitative characteristics in subsurface systems: Basin Research, doi: 10.1111/j.1365-2117.2009.00397.x. Stow, D.A.V., Howell, D.G., and Nelson, C.H., 1985, Sedimentary, tectonic, and sea-level controls, in Bouma, A.H., Normark, W.R., and Barnes, N.E., eds., Submarine Fans and Related Turbidite Systems: New York, Springer-Verlag, p. 15–22. von Huene, R., Ranero, C.R., and Scholl, D.W., 2009, Convergent margin structure in high-quality geophysical images and current kinematic and dynamic models, in Lallemand, S., and Funiciello, F., eds., Subduction Zone Geodynamics: Berlin, Springer-Verlag, p. 137–157. Whipple, K.X., and Tucker, G.E., 1999, Dynamics of the stream power river incision model: implications for height limits of mountain ranges, landscape response timescales and research needs: Journal of Geophysical Research, v. 104, p. 17,661–17,674, doi: 10.1029/1999JB900120. Winker, C.D., and Edwards, M.B., 1983, Unstable progradational clastic shelf margins: SEPM (Society of Sedimentary Geology) Special Publication, v. 33, p. 139–158.

MANUSCRIPT RECEIVED 4 MAY 2010 REVISED MANUSCRIPT RECEIVED 24 OCTOBER 2010 MANUSCRIPT ACCEPTED 17 NOVEMBER 2010

The natural range of submarine canyon-and-channel ...

Pacific to the Atlantic passive margin offshore the Americas (Fig ... overbank deposits across the lower slope and .... from the margin offshore southern California,.

3MB Sizes 1 Downloads 158 Views

Recommend Documents

The natural range of submarine canyon-and ... - GeoScienceWorld
limited to case studies of profiles from a single type of continental margin (e.g., Goff, 2001;. Estrada et al., 2005; Mitchell, 2005; Noda et al.,. 2008; Gerber et al., ...... Aiello, John Damuth, Atsushi Noda, David Piper, and. David Scholl for con

pdf-137\eminent-americans-namesakes-of-the-polaris-submarine ...
... apps below to open or edit this item. pdf-137\eminent-americans-namesakes-of-the-polaris- ... ss-house-document-no-92-345-by-hyman-g-rickover.pdf.

Yellow Submarine ita
YellowSubmarineita.Thesecretseven sisters.C# programming. pdf.Yo gotti 5 ... shall nowinvestigate what happens ifI usea 4x4 grid..068803242952562609.

Idukki Dt -Provisional High Range & Low range Seniority list of ...
Page 2 of 3. '),. -. 47. 3. 6 2012 MinimolChacko VH Chackkupalbm. 48. 6. 6 2012 Linta PA Sub Centre K.Chappathu. 49. 6 2012 Dilip Varqhese Sub Centre Sdnthanpara. 50. 8. 6 2012 sanitha S Nair Sub cent.e Kallar. 51 12. 5 2012 Johney Chacko RPC Kumily.

Consequences of Range Contractions and Range ...
neighboring demes, implying that these edges act as par- tially absorbing ... plus a 5-deme thick layer containing two refuge areas of size 5 В 5 demes. The four gray ..... Page 6 ... The comparison of range shift scenarios with isotropic and anisot

Braiding of submarine channels controlled by aspect ...
Aug 17, 2015 - *e-mail: [email protected]. 700 ..... The wet bulk density of the walnut sand is 1.2 g ml−1 and white plastic sand 1.5 g ml−1. The.

Escape from the Deep∶ The Epic Story of a Legendary Submarine ...
Escape from the Deep∶ The Epic Story of a Legendar ... marine and her Courageous Crew by Alex Kershaw.pdf. Escape from the Deep∶ The Epic Story of a ...

What is the Computational Value of Finite Range Tunneling? (PDF ...
The appeal of. QA relative to SA is ..... Using the weak-strong cluster pairs as building blocks, .... were not computed due to the high computational cost. model ...

The encyclopedia of natural medicine
Rsaggarwalenglish pdf.The glowpt 2.63342512679 - Download Theencyclopedia of ... inked.Sssh koi hai.Windows vista home premiumsp2.Theencyclopedia of ...

The categorization of natural scenes
EOG activity above T 50 AV were excluded from further analysis. ... pass filtering effect of the head as volume conductor—predomi- .... (A) Illustration of the early differential ERP averaged over frontal, centro-parietal and occipito-temporal ...

Morphology and mechanics of submarine spreading: A ...
convex-downslope pattern in plan (Figure 4a). Areas characterized by ...... Resistant Design of Lifeline Facilities and Countermeasures Against Soil. Liquefaction ...

Genetic patterns of a range expansion: The spur ...
providing evidence for spatial and genetic con- gruence. .... tected in the SE region, which could be a con- ... the Spanish Ministry of Education and Science (project: ... Online. 1: 47-. 50. Forlani, A., Crestanello, B., Mantovani, S., Livoreil, B.

Know the range of different business and their ownership.pdf
Notes Chapter 1 - Know the range of different business and their ownership.pdf. Notes Chapter 1 - Know the range of different business and their ownership.pdf.

Watch Wolves of the Range (1921) Full Movie Online Free ...
Watch Wolves of the Range (1921) Full Movie Online Free .MP4__________.pdf. Watch Wolves of the Range (1921) Full Movie Online Free .MP4__________.

Submarine Hydrodynamics (SpringerBriefs in Applied ...
Intended for advanced students and professionals working in the specialised field of submarine hydrodynamics, this book brings theoretical and practical ...

submarine power cables pdf
Whoops! There was a problem loading more pages. Retrying... Whoops! There was a problem previewing this document. Retrying... Download. Connect more ...