New Madrid seismic zone fault geometry - Geology...

12
For permission to copy, contact [email protected] © 2008 Geological Society of America New Madrid seismic zone fault geometry Ryan Csontos Ground Water Institute, University of Memphis, Memphis, Tennessee 38152, USA Roy Van Arsdale Department of Earth Sciences, University of Memphis, Memphis, Tennessee 38152, USA 802 Geosphere; October 2008; v. 4; no. 5; p. 802–813; doi: 10.1130/GES00141.1; 6 figures; 1 animation. ABSTRACT The New Madrid seismic zone of the cen- tral Mississippi River valley has been inter- preted to be a right-lateral strike-slip fault zone with a left stepover restraining bend (Reelfoot reverse fault). This model is overly simplistic because New Madrid seismicity continues 30 km southeast of the stepover. In this study we have analyzed 1704 earth- quake hypocenters obtained between 1995 and 2006 in three-dimensional (3-D) space to more accurately map fault geometry in the New Madrid seismic zone. Most of the earth- quakes appear to align along fault planes. The faults identified include the New Madrid North (29°, 72° SE), Risco (92°, 82° N), Axial (46°, 90°), Reelfoot North (167°, 30° SW), and Reelfoot South (150°, 44° SW) faults. A diffuse zone of earthquakes exists where the Axial fault divides the Reelfoot fault into the Reelfoot North and Reelfoot South faults. Regional mapping of the top of the Precam- brian crystalline basement indicates that the Reelfoot North fault has an average of 500 m and the Reelfoot South fault 1200 m of down-to-the-southwest normal displacement. Since previously published seismic reflec- tion profiles reveal reverse displacement on top of the Paleozoic and younger strata, the Reelfoot North and South faults are herein interpreted to be inverted basement nor- mal faults. The Reelfoot North and Reelfoot South faults differ in strike, dip, depth, and displacement, and only the Reelfoot North fault has a surface scarp (monocline). Thus, the Reelfoot fault is actually composed of two left-stepping restraining bends and two faults that together extend across the entire width of the Reelfoot rift. INTRODUCTION The New Madrid seismic zone, beneath the central Mississippi River valley, is undergoing considerable study as it is the most seismically active area in the eastern United States and also because it experienced at least three devastat- ing earthquakes in 1811–1812 (Nuttli, 1982; Penick, 1981; Johnston, 1996; Johnston and Schweig, 1996; Tuttle et al., 2002) (Fig. 1). This seismicity has been attributed to reactivation of basement faults within the subsurface Cambrian Reelfoot rift (Zoback, 1979; Kane et. al, 1981; Braile et. al, 1986; Thomas, 1989; Dart and Swolfs, 1998). The Reelfoot rift (Mississippi Valley Graben) is a major basement structure mapped from gravity, magnetic, seismic refrac- tion, seismic reflection, and a few deep petro- leum exploration wells (Fig. 2) (Hildenbrand, 1982; Hildenbrand et al., 1982; Hamilton and Zoback, 1982; Howe and Thompson, 1984; Howe, 1985; Hildenbrand and Hendricks, 1995; Dart and Swolfs, 1998; Parrish and Van Arsdale, 2004; Csontos et al., 2008) that is covered by up to 6 km of Phanerozoic sediments. New Madrid seismic zone seismicity is occurring along these rift faults primarily between depths of 4 and 14 km within the Cambrian and Precambrian section (Chiu et al., 1992; Pujol et al., 1997; Van Arsdale and TenBrink, 2000). The seismically most active fault within the New Madrid seismic zone is the Reelfoot reverse fault that extends northwest from near Dyersburg, Tennessee, to New Madrid, Missouri (Figs. 1 and 2) (Van Ars- dale et al., 1995a, 1999). A surface scarp (mono- clinal flexure) due to Reelfoot fault propagation folding (Champion et al., 2001) can be traced from the southwestern margin of Reelfoot Lake, Tennessee, where it is ~3 m high north to a maximum of 10 m where the scarp is truncated by the Mississippi River, to less than 3 m high where the scarp is again truncated by the Mis- sissippi River near New Madrid, Missouri (Van Arsdale et al., 1999) (Fig. 3). Holocene uplift rate on the Reelfoot fault is 1.8 mm/yr (Van Arsdale, 2000), and global positioning system (GPS) data reveal a horizontal convergence rate across the fault of 2.7 mm/yr (Smalley et. al, 2005). These high rates have been explained as Holocene initiation of the New Madrid seismic zone or as a burst of Holocene seismic activity on an old fault (Pratt, 1994; Schweig and Ellis, 1994; Van Arsdale, 2000). In this study we define the geometry of New Madrid seismic zone fault planes by analyz- ing the distribution of earthquake hypocenters as has been done in studies conducted along the San Andreas fault system (e.g., Carena and Suppe, 2004; Pujol et al., 2006) and previously along the Reelfoot fault (Chiu et al., 1992; Liu, 1997; Pujol et al., 1997; Mueller and Pujol, 2001). In this study, however, we use an addi- tional 1704 earthquake hypocenters from 1995 through 2006 and state-of-the-art, 3-D viewing capabilities to better evaluate their distribution. When viewing New Madrid seismic zone earth- quake hypocenters with stereoscopic 3-D soft- ware, it is evident that the New Madrid seismic zone earthquakes occur within discrete zones (fault planes). Delineation of these fault planes provides geometric constraints for future kine- matic analysis of the Reelfoot rift fault system and its earthquakes (e.g., Gomberg and Ellis, 1994; Johnston and Schweig, 1996). As pre- sented below, we also believe that the Reelfoot rift illustrates the association of structural inver- sion and intraplate seismicity that may apply to other intraplate seismic zones. Geology of the New Madrid Seismic Zone Region The southeastern United States is under- lain by several Precambrian terranes welded together to form the North American craton. The Precambrian terrane that underlies the Reelfoot rift is the 1470 ± 30 Ma Eastern Granite Rhyo- lite Province (Atekwana, 1996; Van Schmus et al., 1996, 2007). Drill-hole samples reveal that these rocks consist of granite, granite porphyry, and dioritic gneiss (Thomas, 1988; Dart, 1992; Dart and Swolfs, 1998). The Reelfoot rift (Fig. 2) is interpreted to be a Cambrian intracratonic rift associated with the opening of the Paleozoic Iapetus Ocean (Ervin and McGinnis, 1975; Thomas, 1989, 2006). Reelfoot rifting occurred along northeast-striking

Transcript of New Madrid seismic zone fault geometry - Geology...

Page 1: New Madrid seismic zone fault geometry - Geology …geology-guy.com/environmental/articles/newmadrid_fault...New Madrid seismic zone fault geometry Geosphere, October 2008 803 Figure

For permission to copy, contact [email protected]© 2008 Geological Society of America

New Madrid seismic zone fault geometry

Ryan CsontosGround Water Institute, University of Memphis, Memphis, Tennessee 38152, USA

Roy Van ArsdaleDepartment of Earth Sciences, University of Memphis, Memphis, Tennessee 38152, USA

802

Geosphere; October 2008; v. 4; no. 5; p. 802–813; doi: 10.1130/GES00141.1; 6 fi gures; 1 animation.

ABSTRACT

The New Madrid seismic zone of the cen-tral Mississippi River valley has been inter-preted to be a right-lateral strike-slip fault zone with a left stepover restraining bend (Reelfoot reverse fault). This model is overly simplistic because New Madrid seismicity continues 30 km southeast of the stepover. In this study we have analyzed 1704 earth-quake hypocenters obtained between 1995 and 2006 in three-dimensional (3-D) space to more accurately map fault geometry in the New Madrid seismic zone. Most of the earth-quakes appear to align along fault planes. The faults identifi ed include the New Madrid North (29°, 72° SE), Risco (92°, 82° N), Axial (46°, 90°), Reelfoot North (167°, 30° SW), and Reelfoot South (150°, 44° SW) faults. A diffuse zone of earthquakes exists where the Axial fault divides the Reelfoot fault into the Reelfoot North and Reelfoot South faults. Regional mapping of the top of the Precam-brian crystalline basement indicates that the Reelfoot North fault has an average of 500 m and the Reelfoot South fault 1200 m of down-to-the-southwest normal displacement. Since previously published seismic refl ec-tion profi les reveal reverse displacement on top of the Paleozoic and younger strata, the Reelfoot North and South faults are herein interpreted to be inverted basement nor-mal faults. The Reelfoot North and Reelfoot South faults differ in strike, dip, depth, and displacement, and only the Reelfoot North fault has a surface scarp (monocline). Thus, the Reelfoot fault is actually composed of two left-stepping restraining bends and two faults that together extend across the entire width of the Reelfoot rift.

INTRODUCTION

The New Madrid seismic zone, beneath the central Mississippi River valley, is undergoing considerable study as it is the most seismically

active area in the eastern United States and also because it experienced at least three devastat-ing earthquakes in 1811–1812 (Nuttli, 1982; Penick, 1981; Johnston, 1996; Johnston and Schweig, 1996; Tuttle et al., 2002) (Fig. 1). This seismicity has been attributed to reactivation of basement faults within the subsurface Cambrian Reelfoot rift (Zoback, 1979; Kane et. al, 1981; Braile et. al, 1986; Thomas, 1989; Dart and Swolfs, 1998). The Reelfoot rift (Mississippi Valley Graben) is a major basement structure mapped from gravity, magnetic, seismic refrac-tion, seismic refl ection, and a few deep petro-leum exploration wells (Fig. 2) (Hildenbrand, 1982; Hildenbrand et al., 1982; Hamilton and Zoback, 1982; Howe and Thompson, 1984; Howe, 1985; Hildenbrand and Hendricks, 1995; Dart and Swolfs, 1998; Parrish and Van Arsdale, 2004; Csontos et al., 2008) that is covered by up to 6 km of Phanerozoic sediments. New Madrid seismic zone seismicity is occurring along these rift faults primarily between depths of 4 and 14 km within the Cambrian and Precambrian section (Chiu et al., 1992; Pujol et al., 1997; Van Arsdale and TenBrink, 2000). The seismically most active fault within the New Madrid seismic zone is the Reelfoot reverse fault that extends northwest from near Dyersburg, Tennessee, to New Madrid, Missouri (Figs. 1 and 2) (Van Ars-dale et al., 1995a, 1999). A surface scarp (mono-clinal fl exure) due to Reelfoot fault propagation folding (Champion et al., 2001) can be traced from the southwestern margin of Reelfoot Lake, Tennessee, where it is ~3 m high north to a maximum of 10 m where the scarp is truncated by the Mississippi River, to less than 3 m high where the scarp is again truncated by the Mis-sissippi River near New Madrid, Missouri (Van Arsdale et al., 1999) (Fig. 3). Holocene uplift rate on the Reelfoot fault is 1.8 mm/yr (Van Arsdale, 2000), and global positioning system (GPS) data reveal a horizontal convergence rate across the fault of 2.7 mm/yr (Smalley et. al, 2005). These high rates have been explained as Holocene initiation of the New Madrid seismic zone or as a burst of Holocene seismic activity

on an old fault (Pratt, 1994; Schweig and Ellis, 1994; Van Arsdale, 2000).

In this study we defi ne the geometry of New Madrid seismic zone fault planes by analyz-ing the distribution of earthquake hypocenters as has been done in studies conducted along the San Andreas fault system (e.g., Carena and Suppe, 2004; Pujol et al., 2006) and previously along the Reelfoot fault (Chiu et al., 1992; Liu, 1997; Pujol et al., 1997; Mueller and Pujol, 2001). In this study, however, we use an addi-tional 1704 earthquake hypocenters from 1995 through 2006 and state-of-the-art, 3-D viewing capabilities to better evaluate their distribution. When viewing New Madrid seismic zone earth-quake hypocenters with stereoscopic 3-D soft-ware, it is evident that the New Madrid seismic zone earthquakes occur within discrete zones (fault planes). Delineation of these fault planes provides geometric constraints for future kine-matic analysis of the Reelfoot rift fault system and its earthquakes (e.g., Gomberg and Ellis, 1994; Johnston and Schweig, 1996). As pre-sented below, we also believe that the Reelfoot rift illustrates the association of structural inver-sion and intraplate seismicity that may apply to other intraplate seismic zones.

Geology of the New Madrid Seismic Zone Region

The southeastern United States is under-lain by several Precambrian terranes welded together to form the North American craton. The Precambrian terrane that underlies the Reelfoot rift is the 1470 ± 30 Ma Eastern Granite Rhyo-lite Province (Atekwana, 1996; Van Schmus et al., 1996, 2007). Drill-hole samples reveal that these rocks consist of granite, granite porphyry, and dioritic gneiss (Thomas, 1988; Dart, 1992; Dart and Swolfs, 1998).

The Reelfoot rift (Fig. 2) is interpreted to be a Cambrian intracratonic rift associated with the opening of the Paleozoic Iapetus Ocean (Ervin and McGinnis, 1975; Thomas, 1989, 2006). Reelfoot rifting occurred along northeast- striking

Page 2: New Madrid seismic zone fault geometry - Geology …geology-guy.com/environmental/articles/newmadrid_fault...New Madrid seismic zone fault geometry Geosphere, October 2008 803 Figure

New Madrid seismic zone fault geometry

Geosphere, October 2008 803

Fig

ure

1. R

eelf

oot

rift

bou

nded

by

blac

k lin

es a

nd M

issi

ssip

pi e

mba

ymen

t bo

unde

d by

pur

ple

line.

Mic

roea

rthq

uake

act

ivit

y as

col

ored

dot

s w

ith

star

s at

est

i-m

ated

loca

tion

s of

the

16

Dec

embe

r 18

11 (

sout

h), 2

3 Ja

nuar

y 18

12 (

nort

h), a

nd 7

Feb

ruar

y 18

12 (

cent

ral)

ear

thqu

akes

.

Page 3: New Madrid seismic zone fault geometry - Geology …geology-guy.com/environmental/articles/newmadrid_fault...New Madrid seismic zone fault geometry Geosphere, October 2008 803 Figure

Csontos and Van Arsdale

804 Geosphere, October 2008

-91°

35°

-90°

-90°

-89°

36°

37°

EM

WM

NM

WRFZAF

00 5500 110000 kkmm

RCG

Elevation (m)

Figure 2. Precambrian crystalline basement unconformity structure contour map with basement faults. Reelfoot rift extends from White River Fault Zone (WRFZ) to Reelfoot fault (RF). Contour interval—400 m; elevation relative to sea level. GRTZ—Grand River Tectonic Zone; CMTZ—Central Missouri Tectonic Zone; OFZ—Osceola Fault Zone; BMTZ—Bolivar-Mansfi eld Tectonic Zone; WRFZ—White River Fault Zone; EM—Eastern Rift Margin faults; AF—Axial fault; WM—Western Rift Margin fault; RF—Reelfoot fault; NM—New Madrid; RCG—Rough Creek Graben (from Csontos et al., 2008).

normal faults (Howe and Thompson, 1984; Nelson and Zhang, 1991) that are segmented by northwest-striking vertical faults (Stark, 1997; Csontos et al., 2008). During rifting, the Reelfoot rift accumulated up to 8 km of Cam-brian sediment, while outside the rift 1.5 km of Cambrian sediment accumulated (Howe and Thompson, 1984). Late Cambrian to Middle Ordovician regional subsidence of the rift and surrounding area (Howe and Thompson, 1984; Howe, 1985; Dart and Swolfs, 1998) resulted in deposition of the thick, shallow-marine Ordo-vician Knox (Arbuckle) carbonate Supergroup (Thomas, 1985; Lumsden and Caudle, 2001).

Tectonic landforms within the central Mis-sissippi River valley are directly linked to the underlying Reelfoot rift faults (Fig. 3) (Mihills

and Van Arsdale, 1999; Csontos et al., 2008). Quaternary displacement has been documented along the Eastern Rift Margin (Cox et al., 2001, 2006), Western Rift Margin (Van Arsdale et al., 1995b; Baldwin et al., 2005), Axial (Van Ars-dale, 1998; Guccione et al., 2000), and Reelfoot fault (Russ, 1982; Kelson et al., 1996; Mueller et al., 1999; Van Arsdale et al., 1999; Cham-pion et al., 2001). In addition, the Lake County uplift (essentially coincident with the seismic-ity between the central and northern 1811–1812 earthquakes in Fig. 1), Reelfoot Lake basin (Fig. 3), a segment of the eastern Mississippi Val-ley bluff line (located in westernmost Tennessee between the dark-green uplands and pale-green Mississippi River fl oodplain in Fig. 1), south-ern half of Crowley’s Ridge, the Big Lake and

Lake Saint Francis Sunklands area, and Joiner Ridge are interpreted to be tectonic or tectoni-cally infl uenced landforms (Fig. 4) (Csontos et al., 2008). New Madrid seismic zone faults have also modifi ed Mississippi River gradients and infl uenced sedimentary processes during the Quaternary (Spitz and Schumm, 1997; Guc-cione et al., 2002; Guccione, 2005; Holbrook et al., 2006). Johnston and Schweig (1996) have proposed that the Bootheel lineament produced one of the major earthquakes of the 1811–1812 sequence, and Guccione et al. (2005) report Quaternary displacement on the Bootheel fault (lineament). However, Csontos (2007) mapped the Precambrian basement unconformity and the Pliocene-Pleistocene unconformity surfaces within the Reelfoot rift, and there is no evidence

Page 4: New Madrid seismic zone fault geometry - Geology …geology-guy.com/environmental/articles/newmadrid_fault...New Madrid seismic zone fault geometry Geosphere, October 2008 803 Figure

New Madrid seismic zone fault geometry

Geosphere, October 2008 805

Figure 3. Pliocene-Pleistocene unconformity on top of the Eocene section with vertical projection of Reelfoot rift faults from the top of the Precambrian unconformity. Contour interval—2 m; elevation relative to sea level. B—Big Lake; S—Lake Saint Francis; J—Joiner Ridge; GRTZ—Grand River Tectonic Zone; CMTZ—Central Missouri Tectonic Zone; OFZ—Osceola Fault Zone; BMTZ—Bolivar-Mansfi eld Tectonic Zone; WRFZ—White River Fault Zone; EM—Eastern Rift Margin faults; AF—Axial fault; WM—Western Rift Margin fault; RF—Reelfoot fault; CR—Crowley’s Ridge (from Csontos et al., 2008).

Page 5: New Madrid seismic zone fault geometry - Geology …geology-guy.com/environmental/articles/newmadrid_fault...New Madrid seismic zone fault geometry Geosphere, October 2008 803 Figure

Csontos and Van Arsdale

806 Geosphere, October 2008

Cro

wle

y’s R

idge

Restraining Bends

Uplifted Block

Dropped Block

Possible Subsidence Axis

Earthquake

Figure 4. Deformation of the Pliocene-Pleistocene unconformity surface. Black lines—faults; J—Joiner Ridge; GRTZ—Grand River Tectonic Zone; CMTZ—Central Missouri Tectonic Zone; OFZ—Osceola Fault Zone; BMTZ—Bolivar-Mansfi eld Tectonic Zone; WRFZ—White River Fault Zone; EM—Eastern Rift Margin faults; AF—Axial fault; WM—Western Rift Margin fault; RF—Reel-foot fault. Inset map showing a restraining bend (from Csontos et al., 2008).

Page 6: New Madrid seismic zone fault geometry - Geology …geology-guy.com/environmental/articles/newmadrid_fault...New Madrid seismic zone fault geometry Geosphere, October 2008 803 Figure

New Madrid seismic zone fault geometry

Geosphere, October 2008 807

of vertical displacement or hypocenter align-ment along the Bootheel fault. This lack of ver-tical displacement may be due to a very young Bootheel fault experiencing relatively minor displacement, and/or the displacement is strike-slip. However, in this study we focus on seis-mogenic faults and identifi ed deformation and therefore do not consider a possible aseismic Bootheel fault.

New Madrid Seismic Zone Faults

Many of the Reelfoot rift faults (Fig. 2) have large normal displacements within the crystal-line basement and Cambrian section (Howe 1985; Parrish and Van Arsdale, 2004; Csontos et al., 2008). The Western Rift Margin fault is a steeply eastward-dipping normal fault, displac-ing the Precambrian surface up to 3 km (Howe, 1985; Nelson and Zhang, 1991). The Eastern Rift Margin faults consist of two down-to-the-northwest steep normal faults with as much as 3 km of normal displacement on the western fault, and 0.5 km on the eastern fault (Howe, 1985; Parrish and Van Arsdale, 2004).

Several seismogenic faults have been mapped within the Reelfoot rift (Figs. 1–3). The Axial fault, which trends down the center of the rift, is a near vertical fault (Howe, 1985; Sexton and Jones, 1986; Stephenson et al., 1995; Csontos, 2007). The southwest-dipping Reelfoot reverse fault produces most of the New Madrid seismic zone seismicity (Sexton, 1988; Chiu et al., 1992; Pujol et al., 1997; Mueller and Pujol, 2001). The Reelfoot fault appears to be a segment of the Grand River tectonic zone, and the Reelfoot fault has been interpreted to form the northern margin of the Reelfoot rift (Fig. 2) (Csontos et al., 2008). The third major fault along which earthquakes are occurring is the New Madrid North fault (Baldwin et al., 2005), which is a portion of the Western Rift Margin fault north of New Madrid, Missouri (Fig. 5).

METHODS

The earthquake hypocenters used in this study were obtained from two data sets. One included 568 events recorded between October 1989 and August 1992 using a portable seismic network (PANDA, Chiu et al., 1992). This is the data set used in Mueller and Pujol (2001). The second data set includes 1136 events recorded from 2001 to 2006 by stations of the New Madrid seismic zone permanent network (data available through the Center for Earthquake Research and Informa-tion of the University of Memphis). The earth-quakes in the two data sets were located using the joint hypocentral determination (JHD) tech-nique described in Pujol et al. (1997). This tech-

nique removes the effect of the unconsolidated sediments of the Mississippi embayment, which changes in thickness under the two networks.

This change in thickness and the low P- and S-wave velocities of the post Paleozoic Missis-sippi embayment sediments introduce errors in the earthquake locations when the earthquakes are located individually. The error in depth, in particular, may be as large as 1 km. The errors that may affect JHD locations have been esti-mated by analyzing synthetic data generated for a velocity model that includes the variations in thickness of the unconsolidated sediments (Pujol et al., 1997). Based upon this JHD analy-sis, it was found that the average depth errors are likely to be less than 0.5 km, with the epicentral error considerably smaller.

Earthquake hypocenters were imported into Landmark’s Geoprobe

TM and Schlumberger’s

PetrelTM

for 3-D point rendering using a per-sonal computer (PC) workstation. Active stereo vision in Geoprobe

TM and Petrel

TM was achieved

using Sharper Technology’s CrystalEyes stereo goggles and emitter. The 3-D image was also projected onto a screen for large-format viewing using a DepthQ Stereo3D projector. Earthquake hypocenters appearing to lie on planes (as vis-ible in the 3-D model) were placed into separate data subsets. These subsets were exported to ArcMap

TM and Zmap

TM, and a fi rst-order trend

surface (fault plane) was calculated for each hypocenter subset. In this way, location, strike, dip, and r2 value (a statistical measure of how well the plane fi ts the hypocenter distribution) were calculated for each fault plane in the New Madrid seismic zone. The faults and earthquakes were then imported into a 3-D Geoprobe

TM and

PetrelTM

scene to ensure goodness of fi t to the calculated plane and to determine if there were any outliers. Any obvious hypocenter outliers were removed from the subset, and a fi nal planar trend surface (fault plane) was then created.

To create Reelfoot fault cross sections, the top of the crystalline Precambrian basement was taken from Figure 2, and the top of the Paleo-zoic and Cretaceous sections were obtained from Van Arsdale et al. (1998) and Purser and Van Arsdale (1998).

NEW MADRID SEISMIC ZONE FAULTS

The majority of New Madrid seismic zone seismicity falls along three major trends (Figs. 1 and 5). The longest trend is the 100-km-long southern section, extending northeast along the Axial fault (Chiu et al., 1992; Stephenson et al., 1995). The second trend extends 70 km northwest from near Dyersburg, Tennessee, to New Madrid, Missouri, and corresponds to the Reelfoot fault (Chiu et al., 1992; Pujol

et al., 1997; Van Arsdale et al., 1999; Mueller and Pujol, 2001). The third major seismicity trend, called the New Madrid North fault (John-ston and Schweig, 1996; Baldwin et al., 2005) appears to occur along a segment of the north-western Reelfoot rift margin. A closer look at the earthquake hypocenters reveals that there is also a seismicity trend extending west from New Madrid herein called the Risco fault for the town of Risco, Missouri, and that the Reelfoot fault actually consists of three discrete, seismi-cally defi ned segments (Fig. 5).

Five faults and a diffuse earthquake zone were delineated based upon 3-D earthquake hypo-center trend surface analyses (Animation 1). The Axial fault (r2 = 0.9752) is essentially verti-cal, and so a regression line fi t to the earthquakes defi nes a fault that strikes 46°. The New Madrid North fault strikes 29° and dips 72° SE (r2 = 0.9452), and the Risco fault strikes 92° and dips 82° N (r2 = 0.49). Planes were fi t to the two seg-ments of the Reelfoot fault—the Reelfoot North fault, 167°, 30° SW (r2 = 0.4485), and Reelfoot South fault, 150°, 44° SW (r2 = 0.9121). A diffuse zone of seismicity, with no apparent structure, divides the Reelfoot North and Reelfoot South faults at the intersection with the Axial fault.

DISCUSSION

The Reelfoot Rift and the Reelfoot Fault

Based primarily on gravity and magnetic data (Hildenbrand et al., 1982), the Reelfoot rift steps up (shallows) at the Reelfoot fault and appears to close to the north toward the Rough Creek graben (Fig. 2). This step is interpreted to be the result of Cambrian down-to-the-south-west displacement on the Grand River tectonic zone (Csontos et al., 2008). Since the Reelfoot fault essentially overlies the Grand River tec-tonic zone, we believe that the Reelfoot fault, within the Cambrian and Precambrian section, has down-to-the-southwest normal separation (Fig. 6) (Csontos et al., 2008). However, seis-mic refl ection profi les reveal that the Reelfoot fault displaces the post-Paleozoic section ~70 m in a reverse sense (Sexton, 1988; Van Arsdale et al., 1998; Champion et al., 2001). To allevi-ate this apparent contradiction, we propose that the Reelfoot fault was a rift-margin normal fault during Cambrian extension that was subse-quently inverted into a thrust and reverse fault in late Paleozoic-Appalachian Orogeny and younger compression.

The Axial fault separates the differently ori-ented Reelfoot North and Reelfoot South faults. At the level of the top of the Precambrian, the Reelfoot North fault has 500 m of throw. whereas the Reelfoot South fault has 1200 m of

Page 7: New Madrid seismic zone fault geometry - Geology …geology-guy.com/environmental/articles/newmadrid_fault...New Madrid seismic zone fault geometry Geosphere, October 2008 803 Figure

Csontos and Van Arsdale

808 Geosphere, October 2008

A Fig

ure

5 (c

ontin

ued

on n

ext p

age)

. (A

) New

Mad

rid

seis

mic

zon

e se

ism

icit

y w

ith

faul

ts m

appe

d by

Ste

phen

son

et a

l. (1

995)

wit

hin

the

zone

of d

iffu

se s

eism

ic-

ity.

A–A

' is

line

of c

ross

sec

tion

for

B. C

ross

-sec

tion

B–B

' and

C–C

' loc

ated

in F

igur

e 6.

Page 8: New Madrid seismic zone fault geometry - Geology …geology-guy.com/environmental/articles/newmadrid_fault...New Madrid seismic zone fault geometry Geosphere, October 2008 803 Figure

New Madrid seismic zone fault geometry

Geosphere, October 2008 809

'

B

Axial Fault

UnnamedFaultCottonwoodGrove Fault

Ridgely RidgeFault

Fig

ure

5 (c

ontin

ued)

. (B

) N

orth

wes

t-so

uthe

ast A

–A’ R

eelf

oot

thru

st f

ault

s sh

owin

g ea

rthq

uake

foc

i. R

eelf

oot

faul

t se

gmen

ts—

oran

ge a

nd b

lue;

Axi

al f

ault

—re

d; a

nd t

he a

rcua

te

diff

use

zone

—bl

ack.

Das

hed

lines

den

ote

prim

ary

dept

h to

sei

smic

ity

coin

cide

nt w

ith

dept

h to

cha

nge

in d

ip o

n F

igur

e 6.

Page 9: New Madrid seismic zone fault geometry - Geology …geology-guy.com/environmental/articles/newmadrid_fault...New Madrid seismic zone fault geometry Geosphere, October 2008 803 Figure

Csontos and Van Arsdale

810 Geosphere, October 2008

throw (Fig. 6). Thus, the Axial fault experienced down-to-the-southeast displacement during Cam-brian rifting (Csontos, 2007).

New Madrid Seismic Zone Earthquakes

Trend surface analyses of earthquake hypo-centers illustrate that most earthquakes within the New Madrid seismic zone occur along well-defi ned fault planes (Figs. 3–5, Animation 1). Specifi cally, earthquakes defi ne the Reelfoot North, Reelfoot South, Axial, Risco, and New Madrid North faults.

The deep, seismically active portion of the Reelfoot fault is divided into the Reelfoot North fault that dips 30° SW and the Reelfoot South fault that dips 44° SW (Fig. 6). However, seis-mic refl ection data reveal that the Reelfoot fault dips 73° (Van Arsdale et al., 1998) to perhaps as much as 80° (Champion et al., 2001) at depths less than 1 km, and the change in fault dip has been placed at the top of the Precambrian crystalline basement (Purser and Van Arsdale, 1998). This fault dip change was estimated by Purser and Van Arsdale (1998) to be at 4 km depth, which compares reasonably well with the footwall elevation of the Precambrian surface defi ned in Figure 6. The elevation of the top of the Precambrian in the hanging wall of the Reel-foot normal fault is −3700 m across the Reelfoot North fault and −4800 m across the Reelfoot South fault. It is also apparent that the minimum depth of seismicity on the Reelfoot South fault is deeper than on the Reelfoot North fault (Fig. 5). In both of these faults the seismicity is occur-ring on the deeper shallow-dipping portion of the fault planes. An additional major difference

between the Reelfoot North and Reelfoot South faults is that there is no surface scarp along the Reelfoot South fault (Van Arsdale et al., 1999). The relatively low regression (r2 = 0.4485) value for the Reelfoot North fault plane has at least two possible explanations. The Reelfoot North fault may be slightly curved (Animation 1), which would show a poor fi t to a single planar

surface. Alternatively, the Reelfoot North fault may be more complex than presented and may contain more than one fault plane. In summary, the Reelfoot North and Reelfoot South faults differ in strike, dip, basement displacement, sur-face deformation, and earthquake depths.

The seismicity of the Axial fault, where it divides the Northern and Southern Reelfoot

Animation 1. Three-dimensional model of New Madrid seismic zone with best-fi t fault planes. You will need Adobe Acro-bat or Adobe Reader Version 8 or later to view and rotate this fi le. If you are viewing the PDF of this paper or reading it offl ine, please visit http://dx.doi.org/10.1130/GES00141.S1 (Animation 1) or the full-text article on www.gsajournals.org to view Animation 1.

B Cretaceous TopPaleozoic Unconformity

Precambrian Unconformity

Reelfoot Fault

-430m

-530m

B'

-3200m

500m

0 500 1000m

80˚

30˚

44˚

1200m

Cretaceous TopPaleozoic Unconformity 80˚

-430m

-530m

-3600m

Precambrian Unconformity

0 500 1000mReelfo

ot Fa

ult

C C'

Figure 6. Generalized cross-section B–B' of the inverted Reelfoot fault across the Reelfoot North fault segment. Generalized cross-section C–C' of the inverted Reelfoot fault across the Reelfoot South fault segment. Cross-section scales are 1:1 and are located in Figure 5. Minor hanging-wall deformation is not shown due to scale.

Page 10: New Madrid seismic zone fault geometry - Geology …geology-guy.com/environmental/articles/newmadrid_fault...New Madrid seismic zone fault geometry Geosphere, October 2008 803 Figure

New Madrid seismic zone fault geometry

Geosphere, October 2008 811

faults, is a structureless cloud of earthquakes (Fig. 5). The only seismicity pattern within this zone is a 15-km-wide arched distribution around the intersection of the Axial and Reelfoot faults. This cloud of seismicity corresponds with the northeast-striking vertical basement Axial fault, which locally includes the Unnamed, Cotton-wood Grove, and Ridgely faults mapped within the overlying Phanerozoic section (Stephenson et al., 1995; Van Arsdale et al., 1998), appar-ently refl ecting a wide zone of distributed shear (Odum et al., 1998). However, this zone of structureless seismicity is wider than these four faults, thus suggesting that additional parallel right-lateral strike-slip faults may exist south-east of the Ridgely Ridge fault and northwest of the Axial fault within this zone (Fig. 5).

The Risco fault was defi ned by fi tting a plane to the earthquake hypocenters west of New Madrid (Fig. 5). Based on these hypocenter locations, this fault strikes 92° and dips 82° N. Fault plane solutions (Johnston and Schweig, 1996) reveal left-lateral strike-slip motion on the Risco fault.

CONCLUSIONS

New Madrid seismic zone seismicity is occurring along Cambrian fault planes of the Reelfoot rift. Below ~4 km, the Reelfoot fault is interpreted to be the reactivation of a Cam-brian down-to-the-southwest normal fault that may form the northern boundary of the Reelfoot rift. More specifi cally, the Reelfoot North, Reel-foot South, and Axial faults bound Reelfoot rift subbasins (Fig. 2). Earthquakes are occurring primarily between 4 and 14 km along the 30°-dipping Reelfoot North and 44°-dipping Reel-foot South faults. However, above 4 km, both of these faults dip ~80°, are largely aseismic, and have reverse displacement. We believe that these faults at depth are normal faults that have been structurally inverted probably initially during Paleozoic Appalachian-Ouachita Orogeny con-tinental compression.

The New Madrid seismic zone has been explained to be a right-lateral shear zone with a compressional left stepover (Russ, 1982; Sch-weig and Ellis, 1994). The faults within the current model are the seismogenic right-lateral Axial fault, the compressional Reelfoot reverse fault (our Reelfoot North fault), and the right-lateral New Madrid North fault. This model requires that crustal rock is sliding northeast along the Axial fault and being forced over and around the Reelfoot North fault to continue slid-ing northeast along the New Madrid North fault. Indeed, in trench excavations across the Reelfoot North fault scarp, Kelson et al. (1996) cite evi-dence of right-lateral slip. This is important for

two reasons. If there is a signifi cant component of strike-slip on the Reelfoot North fault, then the ~1.8 m of coseismic reverse slip (Merritts and Hesterberg, 1994; Mueller and Pujol, 2001) determined for the 7 February 1812 Reelfoot fault earthquake may be only a component of greater oblique slip. Secondly, this “crowding” around the Reelfoot North fault may be respon-sible for the Risco fault. We believe that the left-lateral Risco fault is an escape fault caused by this crowding around the Reelfoot North fault stepover (Csontos, 2007).

Right-lateral shear on the New Madrid North and Axial faults explains the Reelfoot North fault (Odum et al., 1998) but does not explain the seis-micity occurring southeast of the Axial fault on the Reelfoot South fault. It is necessary to modify the current model for the New Madrid seismic zone. We propose that the Reelfoot South fault is also a compressional left stepover between the right-lateral Axial fault and the right-lateral East-ern Rift Margin faults (Cox et al., 2001, 2006; Csontos et al., 2008). This is supported by the right-lateral fault plane solutions for the Axial and Eastern Rift Margin faults and the reverse fault plane solutions for the Reelfoot South fault (Chiu et al., 1992, 1997; Liu, 1997). The contem-porary stress fi eld for the eastern United States is a horizontal maximum compressive stress oriented N60°E with some central Mississippi River valley measurements indicating due east (Zoback and Zoback, 1989; Zoback, 1992; Ellis, 1994). This stress fi eld, and resultant right-lateral shear on the N45°E-oriented Reelfoot rift faults, has led numerous authors to invoke right-lateral shear as being responsible for the New Madrid seismic zone. However, we believe that contem-porary right-lateral shear is not restricted to the Axial–Reelfoot North–New Madrid North fault system, but is occurring over the full width of the Reelfoot rift at its northeastern end (Odum et al., 1998; Csontos et al., 2008). We also believe that recent GPS data support these conclusions (Smalley et al., 2005).

Perhaps most diffi cult to explain in our pro-posed New Madrid seismic zone shear model are parts that at fi rst glance appear to be missing. For example, the southeastern Reelfoot rift mar-gin is only moderately active, but commensurate with our model, the portion north of Memphis is most active (Fig. 1). If the Axial fault is act-ing as a right-lateral fault for the Reelfoot South fault stepover, then the Axial fault should extend northeast of the Reelfoot faults. Van Arsdale et al. (1998) have proposed that the Cottonwood Grove and Ridgely faults may extend northeast of the Reelfoot fault (Fig. 5), and Figures 4–5 and Animation 1 suggest right-lateral displace-ment of the Reelfoot North fault with respect to the Reelfoot South fault. Our model also sug-

gests that there should be earthquakes along a northeastern continuation of the Axial fault beyond the Reelfoot faults. To this we can only speculate that the northeastern extension of the Axial fault is currently aseismic or that the dis-placement is occurring as reverse movement on the Reelfoot North fault.

Van Arsdale et al. (1999) and Mueller and Pujol (2001) argue that the Reelfoot fault is continuous from New Madrid, Missouri, south to near Dyersburg, Tennessee. However, our research indicates that the Reelfoot North and Reelfoot South are discrete fault segments that may not rupture as one fault. These faults differ in strike, dip, depth, area, basement displace-ment, and the Reelfoot South fault does not have a surface scarp. Absence of a fault scarp along the Reelfoot South fault could be due to the fact that this fault is east of the Mississippi River fl oodplain and to displace the ground surface the fault would have to pass through an addi-tional 30 m of unconsolidated sediment includ-ing a thick loess deposit. Although this may be an important factor as to why there is no scarp along the Reelfoot South fault, we believe this is not the primary reason.

The Reelfoot North fault scarp (fault propa-gation fold) is symmetric in that it diminishes in structural amplitude north and south (Van Arsdale et al., 1999) from its maximum fold amplitude of 11 m to the northwest of Reelfoot Lake (Champion et al., 2001). The fold ampli-tude is less than 3 m within the large Missis-sippi River bend (Kentucky Bend) area of Fig-ure 3 and at its southern end where it intersects the Axial fault (Van Arsdale et al., 1999). Van Arsdale et al. (1999) argue that the southern end of the Reelfoot scarp may have been ero-sionally removed by Mississippi River fl ood waters prior to the construction of the Missis-sippi River levees. Although we agree with this, when we extrapolate a diminishing fold ampli-tude (scarp height) south from its maximum of 11 m, the fold would diminish to near zero at the eastern edge of the Mississippi River fl ood-plain. If the Reelfoot fault has ruptured over its entire 70 km of its mapped bedrock length as one fault plane during its Holocene history, we would not expect the fold amplitude (scarp height) to diminish to zero at the southern end of the Reelfoot North fault. Quite the contrary: if the Reelfoot fault is truly continuous and has ruptured along its entire 70 km length during the late Quaternary, we would expect its fault prop-agation fold amplitude to be a maximum near its center (Van der Pluijm and Marshak, 2004) where the Reelfoot North and Reelfoot South faults meet.

Van Arsdale et al. (1999) document subtle tectonic geomorphic features along the Reelfoot

Page 11: New Madrid seismic zone fault geometry - Geology …geology-guy.com/environmental/articles/newmadrid_fault...New Madrid seismic zone fault geometry Geosphere, October 2008 803 Figure

Csontos and Van Arsdale

812 Geosphere, October 2008

South fault and the short-lived impoundment of the Obion River during the 12 February 1812 New Madrid earthquake that may have been caused by fault displacement. However, based on the above arguments, we believe that the paleoseismic history documented along the Reelfoot North fault (Kelson et al., 1996; Muel-ler et al., 1999) probably does not apply to the Reelfoot South fault. We close by speculating that the Holocene earthquake history of the Reelfoot South fault may have begun with the 12 February 1812 earthquake and that ongoing seismicity reveals that this southeastern stepover zone is an area of strain accumulation (Cox et al., 2001, 2006).

ACKNOWLEDGMENTS

We wish to thank the National Science Founda-tion Mid America Earthquake Center, Memphis Stone and Gravel, and the University of Memphis Dunavant Grant for funding this research. We also wish to thank Landmark Graphics Company and Schlumberger for their contribution of software and technical support, and the North American Coal Company for their bore-hole logs. We also acknowledge Jose Pujol for his contribution of earthquake hypocenter data and exper-tise and comments by Randel Cox.

REFERENCES CITED

Atekwana, E.A., 1996, Precambrian basement beneath the central Midcontinent United States as interpreted from potential fi eld imagery: Basement of eastern North America: Geological Society of America Special Paper, v. 308, p. 33–44.

Baldwin, J.N., Harris, J.B., Van Arsdale, R.B., Givler, R., Kelson, K.I., Sexton, J.L., and Lake, M., 2005, Con-straints on the location of the Late Quaternary Reel-foot and New Madrid North faults in the northern New Madrid seismic zone, central United States: Seismo-logical Research Letters, v. 76, p. 772–789.

Braile, L.W., Hinze, W.J., Keller, G.R., Lidiak, E.G., and Sexton, J.L., 1986, Tectonic development of the New Madrid complex, Mississippi embayment, North America: Tectonophysics, v. 131, p. 1–21, doi: 10.1016/0040-1951(86)90265-9.

Carena, S., and Suppe, J., 2004, Lack of continuity of the San Andreas Fault in southern California: Three-dimensional fault models and earthquake scenarios: Journal of Geophysical Research, v. 109, p. B04313, doi: 10.1029/2003JB002643.

Champion, J., Mueller, K., Tate, A., and Guccione, M., 2001, Geometry, numerical models and revised slip rate for the Reelfoot fault and trishear fault-propagation fold, New Madrid seismic zone: Engineering Geology, v. 62, p. 31–49, doi: 10.1016/S0013-7952(01)00048-5.

Chiu, J.M., Johnston, A.C., and Yang, Y.T., 1992, Imaging the active faults of the central New Madrid seismic zone using PANDA array data: Seismological Research Letters, v. 63, p. 375–393.

Chiu, S.C., Chiu, J.M., and Johnston, A.C., 1997, Seismicity of the southeastern margin of Reelfoot rift, central United States: Seismological Research Letters, v. 68, p. 785–794.

Cox, R.T., Van Arsdale, R.B., Harris, J.B., and Larsen, D., 2001, Neotectonics of the southeastern Reelfoot rift zone margin, central United States, and implications for regional strain accommodation: Geology, v. 29, p. 419–422, doi: 10.1130/0091-7613(2001)029<0419:NOTSRR>2.0.CO;2.

Cox, R.T., Cherryhomes, J., Harris, J.B., Larsen, D., Van Arsdale, R.B., and Forman, S.L., 2006, Paleoseismol-ogy of the southeastern Reelfoot rift in western Tennes-see and implications for intraplate fault zone evolution: Tectonics, v. 23, p. 1–17.

Csontos, R., Van Arsdale, R., Cox, R., and Waldron, B., 2008, Reelfoot rift and its impact on Quaternary deformation in the central Mississippi River valley: Geosphere, v. 4, p. 145–158, doi: 10.1130/GES00107.1.

Csontos, R.M., 2007, Three dimensional modeling of the Reel-foot rift and New Madrid seismic zone [Ph.D. thesis]: Memphis, Tennessee, University of Memphis, 92 p.

Dart, R.L., 1992, Catalog of pre-Cretaceous geologic drill-hole data from the upper Mississippi embayment: A revision and update of Open-File Report 90-260: U.S. Geological Survey Open-File Report, 253 p.

Dart, R.L., and Swolfs, H.S., 1998, Contour mapping of relic structures in the Precambrian basement of the Reelfoot rift, North American Midcontinent: Tectonics, v. 17, p. 235–249, doi: 10.1029/97TC03551.

Ellis, W.L., 1994, Summary and discussion of crustal stress data in the region of the New Madrid seismic zone: U.S. Geo-logical Survey Professional Paper 1538-A–C, p. B1–B13.

Ervin, C.P., and McGinnis, L.D., 1975, Reelfoot rift: Reac-tivated precursor to the Mississippi embayment: Geo-logical Society of America Bulletin, v. 86, p. 1287–1295, doi: 10.1130/0016-7606(1975)86<1287:RRRPTT>2.0.CO;2.

Gomberg, J., and Ellis, M., 1994, Topography and tec-tonics of the central New Madrid seismic zone: Results of numerical experiments using a three-dimensional boundary element program: Journal of Geophysical Research, v. 99, p. 20,299–20,310, doi: 10.1029/94JB00039.

Guccione, M.J., 2005, Late Pleistocene and Holocene paleo-seismology of an intraplate seismic zone in a large allu-vial valley, the New Madrid seismic zone, central USA: Tectonophysics, v. 408, p. 237–264.

Guccione, M.J., Van Arsdale, R.B., and Lynne, H.H., 2000, Origin and age of the Manilla high and associated Big Lake “sunklands” in the New Madrid seismic zone, northeastern Arkansas: Geological Society of America Bulletin, v. 112, p. 579–590, doi: 10.1130/0016-7606(2000)112<0579:OAAOTM>2.3.CO;2.

Guccione, M.J., Mueller, K., Champion, J., Shepherd, S., Carlson, S.D., Odhiambo, B., and Tate, A., 2002, Stream response to repeated coseismic folding, Tiptonville dome, New Madrid seismic zone: Geo-morphology, v. 43, p. 313–349, doi: 10.1016/S0169-555X(01)00145-3.

Guccione, M.J., Marple, R., and Autin, W.J., 2005, Evidence for Holocene displacements on the Bootheel fault (lineament) in southeastern Missouri: Seismotectonic implications for the New Madrid region: Geological Society of America Bulletin, v. 117, p. 319–333, doi: 10.1130/B25435.1.

Hamilton, R.M., and Zoback, M.D., 1982, Tectonic features of the New Madrid seismic zone from refl ection profi les, in McKeown, F.A., and Pakiser, L.C., eds., Investigations of the New Madrid, Missouri, earthquake region: U.S. Geo-logical Survey Professional Paper 1236-F, p. 55–82.

Hildenbrand, T.G., 1982, Model of the southeastern margin of the Mississippi Valley graben near Memphis, Ten-nessee, from interpretation of truck-magnetometer data: Geology, v. 10, p. 476–480, doi: 10.1130/0091-7613(1982)10<476:MOTSMO>2.0.CO;2.

Hildenbrand, T.G., and Hendricks, J.D., 1995, Geophysical setting of the Reelfoot rift and relation between rift structures and the New Madrid seismic zone, in Shed-lock, K.M., and Johnson, A.C., eds., Investigations of the New Madrid seismic zone: U.S. Geological Survey Professional Paper 1538-E, p. 1–30.

Hildenbrand, T.G., Kane, M.F., and Hendricks, J.D., 1982, Magnetic basement in the upper Mississippi embay-ment region—A preliminary report, in McKeown, F.A., and Pakiser, L.C., eds., Investigations of the New Madrid, Missouri, earthquake region: U.S. Geological Survey Professional Paper 1236-E, p. 39–53.

Holbrook, J., Whitney, J.A., Rittenour, T.M., Marshak, S., and Goble, R.J., 2006, Stratigraphic evidence for mil-lennial-scale temporal clustering of earthquakes on a continental-interior fault: Holocene Mississippi River fl oodplain deposits, New Madrid seismic zone, USA: Tectonophysics, v. 420, p. 431–454, doi: 10.1016/j.tecto.2006.04.002.

Howe, J.R., 1985, Tectonics, sedimentation, and hydrocar-bon potential of the Reelfoot aulacogen [M.S. thesis]: Norman, University of Oklahoma, 109 p.

Howe, J.R., and Thompson, T.L., 1984, Tectonics, sedimen-tation, and hydrocarbon potential of the Reelfoot rift: Oil and Gas Journal, v. 82, p. 179–190.

Johnston, A.C., 1996, Seismic moment assessment of sta-ble continental earthquakes, Part 3: 1811–1812 New Madrid, 1886 Charleston, and 1755 Lisbon: Geo-physical Journal International, v. 126, p. 314–344, doi: 10.1111/j.1365-246X.1996.tb05294.x.

Johnston, A.C., and Schweig, E.S., 1996, The enigma of the New Madrid earthquakes of 1811–1812: Annual Review of Earth and Planetary Sciences, v. 24, p. 339–384, doi: 10.1146/annurev.earth.24.1.339.

Kane, M.F., Hildenbrand, T.G., and Hendricks, J.D., 1981, Model for the tectonic evolution of the Mississippi embayment and its contemporary seismicity: Geology, v. 9, p. 563–568, doi: 10.1130/0091-7613(1981)9<563:MFTTEO>2.0.CO;2.

Kelson, K.I., Simpson, G.D., Van Arsdale, R.B., Harris, J.B., Haraden, C.C., and Lettis, W.R., 1996, Mul-tiple Late Holocene earthquakes along the Reelfoot fault, central New Madrid seismic zone: Journal of Geophysical Research, v. 101, p. 6151–6170, doi: 10.1029/95JB01815.

Liu, Z., 1997, Earthquake modeling and active faulting in the New Madrid seismic zone [Ph.D. thesis]: St. Louis, Missouri, St. Louis University, 164 p.

Lumsden, D.N., and Caudle, G.C., 2001, Origin of massive dolostone: The upper Knox model: Journal of Sedimen-tary Research, v. 71, p. 400–409, doi: 10.1306/2DC4094E-0E47-11D7-8643000102C1865D.

Merritts, D., and Hesterberg, T., 1994, Stream networks and long-term surface uplift in the New Madrid seismic zone: Science, v. 265, p. 1081–1084, doi: 10.1126/sci-ence.265.5175.1081.

Mihills, R.K., and Van Arsdale, R.B., 1999, Late Wiscon-sin to Holocene New Madrid seismic zone deforma-tion: Seismological Society of America Bulletin, v. 89, p. 1019–1024.

Mueller, K., and Pujol, J., 2001, Three-dimensional geometry of the Reelfoot blind thrust: Implications for moment release and earthquake magnitude in the New Madrid seismic zone: Seismological Society of America Bulle-tin, v. 91, p. 1563–1573, doi: 10.1785/0120000276.

Mueller, K., Champion, J., Guccione, M., and Kelson, K., 1999, Fault slip rates in the modern New Madrid seis-mic zone: Science, v. 286, p. 1135–1138, doi: 10.1126/science.286.5442.1135.

Nelson, K.D., and Zhang, J., 1991, A COCORP deep refl ec-tion profi le across the buried Reelfoot rift, south-cen-tral United States: Tectonophysics, v. 197, p. 271–293, doi: 10.1016/0040-1951(91)90046-U.

Nuttli, O.W., 1982, Damaging earthquakes of the central Mississippi Valley, in McKeown, F.A., and Pakiser, L.C., eds., Investigations of the New Madrid, Missouri, earthquake region: U.S. Geological Survey Profes-sional Paper 1236-B, p. 15–20.

Odum, J.K., Stephenson, W.J., and Shedlock, K.M., 1998, Near-surface structural model for deformation associ-ated with the February 7, 1812 New Madrid, Missouri, earthquake: Geological Society of America Bulletin, v. 110, p. 149–162, doi: 10.1130/0016-7606(1998)110<0149:NSSMFD>2.3.CO;2.

Parrish, S., and Van Arsdale, R., 2004, Faulting along the southeastern margin of the Reelfoot rift in northwestern Tennessee revealed in deep seismic refl ection profi les: Seismological Research Letters, v. 75, p. 782–791.

Penick, J.L., 1981, The New Madrid earthquakes: Columbia, University of Missouri Press, 176 p.

Pratt, T.L., 1994, How old is the New Madrid seismic zone?: Seismological Research Letters, v. 65, p. 172–179.

Pujol, J., Johnston, A., Chiu, J., and Yang, Y., 1997, Refi ne-ment of thrust faulting models for the central New Madrid seismic zone: Engineering Geology, v. 46, p. 281–298, doi: 10.1016/S0013-7952(97)00007-0.

Pujol, J., Mueller, K., Shen, P., and Chitupolu, V., 2006, High-resolution 3D P-wave velocity model for the East Ven-tura-San Fernando Basin, California, and relocation of events in the Northridge and San Fernando aftershock sequences: Seismological Society of America Bulletin, v. 96, p. 2269–2280, doi: 10.1785/0120050237.

Purser, J.L., and Van Arsdale, R.B., 1998, Structure of the Lake County uplift: New Madrid seismic zone: Seismological Society of America Bulletin, v. 88, p. 1204–1211.

Page 12: New Madrid seismic zone fault geometry - Geology …geology-guy.com/environmental/articles/newmadrid_fault...New Madrid seismic zone fault geometry Geosphere, October 2008 803 Figure

New Madrid seismic zone fault geometry

Geosphere, October 2008 813

Russ, D.P., 1982, Style and signifi cance of surface deformation in the vicinity of New Madrid, Missouri: Investigations of the New Madrid, Missouri, earthquake region: U.S. Geo-logical Survey Professional Paper 1236, p. 95–114.

Schweig, E., and Ellis, M.A., 1994, Reconciling short recur-rence intervals with minor deformation in the New Madrid seismic zone: Science, v. 264, p. 1308–1311, doi: 10.1126/science.264.5163.1308.

Sexton, J.L., 1988, Seismic refl ection expression of reacti-vated structures in the New Madrid rift complex: Seis-mological Research Letters, v. 59, p. 141–150.

Sexton, J.L., and Jones, P.B., 1986, Evidence for recurrent faulting in the New Madrid seismic zone from mini-sosie high-resolution refl ection data: Geophysics, v. 51, p. 1760–1788, doi: 10.1190/1.1442224.

Smalley, R., Jr., Ellis, M.A., Paul, J., and Van Arsdale, R., 2005, Space geodetic evidence for rapid strain rates in the New Madrid seismic zone of central USA: Nature, v. 435, p. 1088–1090, doi: 10.1038/nature03642.

Spitz, W.J., and Schumm, S.A., 1997, Tectonic geomor-phology of the Mississippi Valley between Osceola, Arkansas and Friars Point, Mississippi: Engineering Geology, v. 46, p. 259–280, doi: 10.1016/S0013-7952(97)00006-9.

Stark, J.T., 1997, The East Continent rift complex: Evidence and conclusions: Geological Society of America Spe-cial Paper 312, p. 253–266.

Stephenson, W.J., Shedlock, K.M., and Odum, J.K., 1995, Characterization of the Cottonwood Grove and Rid-gely faults near Reelfoot Lake, Tennessee, from high-resolution seismic refl ection data: U.S. Geological Survey Professional Paper 1538-I, p. 1–10.

Thomas, W.A., 1985, The Appalachian-Ouachita connec-tion: Paleozoic orogenic belt at the southern margin of North America: Annual Review of Earth and Planetary Sciences, v. 13, p. 175–199, doi: 10.1146/annurev.ea.13.050185.001135.

Thomas, W.A., 1988, The Black Warrior Basin, in Sloss, L.L., ed., Sedimentary cover—North American craton:

Boulder, Colorado, Geological Society of America, The Geology of North America, v. D-2, p. 471–492.

Thomas, W.A., 1989, The Appalachian-Ouachita orogen beneath the Gulf Coastal Plain between the outcrops in the Appalachian and Ouachita Mountains, in Hatcher, R.D., Thomas, W.A., and Viele, G.W., eds., The Appa-lachian-Ouachita Orogen in the United States: Boulder, Colorado, Geological Society of America, Geology of North America, v. F-2, p. 537–553.

Thomas, W.A., 2006, Tectonic inheritance at a continental margin: GSA Today, v. 16, no. 2, p. 4–11, doi: 10.1130/1052-5173(2006)016[4:TIAACM]2.0.CO;2.

Tuttle, M.P., Schweig, E.S., Sims, J.D., Lafferty, R.H., Wolf, L.W., and Haynes, M.L., 2002, The earthquake poten-tial of the New Madrid seismic zone: Seismological Society of America Bulletin, v. 92, p. 2080–2089, doi: 10.1785/0120010227.

Van Arsdale, R.B., 1998, Seismic hazards of the upper Missis-sippi embayment: U.S. Army Corps of Engineers Water-ways Experiment Station contract report GL-98-1, 126 p.

Van Arsdale, R.B., 2000, Displacement history and slip rate on the Reelfoot fault of the New Madrid seismic zone: Engineering Geology, v. 55, p. 219–226, doi: 10.1016/S0013-7952(99)00093-9.

Van Arsdale, R.B., and TenBrink, R., 2000, Late Cretaceous and Cenozoic geology of the New Madrid seismic zone: Seismological Society of America Bulletin, v. 90, p. 345–356, doi: 10.1785/0119990088.

Van Arsdale, R.B., Kelson, K.I., and Lumsden, C.H., 1995a, Northern extension of the Tennessee Reelfoot scarp into Kentucky and Missouri: Seismological Research Letters, v. 55, p. 57–62.

Van Arsdale, R.B., Williams, R.A., Schweig, E.S., Shedlock, K.M., Odum, J.K., and King, K.W., 1995b, The origin of Crowley’s Ridge, northeastern Arkansas: Erosional remnant or tectonic uplift?: Seismological Society of America Bulletin, v. 85, p. 963–986.

Van Arsdale, R.B., Purser, J.L., Stephenson, W., and Odum, J., 1998, Faulting along the southern margin of Reel-

foot Lake, Tennessee: Seismological Society of Amer-ica Bulletin, v. 88, p. 131–139.

Van Arsdale, R.B., Cox, R.T., Johnston, A.C., Stephenson, W.J., and Odum, J.K., 1999, Southeastern extension of the Reelfoot fault: Seismological Research Letters, v. 70, p. 348–359.

Van der Pluijm, B.A., and Marshak, S., 2004, Earth struc-ture: An introduction to structural geology and tecton-ics: New York, W.W. Norton and Company, 656 p.

Van Schmus, W.R., Bickford, M.E., and Turek, A., 1996, Proterozoic geology of the east-central midcontinent basement: Basement and basins of eastern North America: Geological Society of America Special Paper 308, p. 7–32.

Van Schmus, W.R., Schneider, D.A., Holm, K.D., Dodson, S., and Nelson, B.K., 2007, New insights into the south-ern margin of the Archean-Proterozoic boundary in the north-central United States based on U-Pb, Sm-Nd, and Ar-Ar geochronology: Precambrian Research, v. 157, p. 80–105, doi: 10.1016/j.precamres.2007.02.011.

Zoback, M.D., 1979, Recurrent faulting in the vicinity of Reelfoot Lake, northwestern Tennessee: Geo-logical Society of America Bulletin, v. 90, p. 1019–1024, doi: 10.1130/0016-7606(1979)90<1019:RFITVO>2.0.CO;2.

Zoback, M.L., 1992, Stress fi eld constraints on intraplate seismicity in eastern North America: Journal of Geophysical Research, v. 97, p. 11761–11782, doi: 10.1029/92JB00221.

Zoback, M.L., and Zoback, M.D., 1989, Tectonic stress fi eld of the continental United States, in Pakiser, L., and Mooney, W., eds., Geophysical framework of the con-tinental United States: Geological Society of America Memoir 172, p. 523–540.

MANUSCRIPT RECEIVED 13 JULY 2007REVISED MANUSCRIPT RECEIVED 19 MAY 2008MANUSCRIPT ACCEPTED 27 JUNE 2008