+ All Categories
Transcript
Page 1: Testing sequence stratigraphic models by drilling Miocene … · 2013-12-10 · Drilling foresets and testing sequence stratigraphic models Geosphere, October 2013 1237 confusion

1236

Testing sequence stratigraphic models by drilling Miocene foresets on the New Jersey shallow shelf

Kenneth G. Miller1, Gregory S. Mountain1, James V. Browning1, Miriam E. Katz2, Donald Monteverde1, Peter J. Sugarman1, Hisao Ando3, Maria A. Bassetti4, Christian J. Bjerrum5, David Hodgson6, Stephen Hesselbo7, Sarp Karakaya1, Jean-Noel Proust,8 and Marina Rabineau9

1Department of Earth and Planetary Sciences, Rutgers University, Piscataway, New Jersey 08854, USA2Earth & Environmental Sciences, Rensselaer Polytechnic Institute, 1W08 JRSC, Troy, New York 12180, USA3Department of Earth Sciences, Faculty of Science, Ibaraki University, Bunkyo 2-1-1, Mito 310-8512, Japan4Laboratoire CEFREM Bat U, University of Perpignan, 52 Avenue Paul Alduy, Perpignan, 66860, France5Department of Geography and Geology, University of Copenhagen, Øster Voldgade 10, DK-1350 København K, Denmark6School of Earth and Environment, University of Leeds, Woodhouse Lane, Leeds LS2 9JT, UK7Department of Earth Sciences, University of Oxford, South Parks Road, Oxford OX1 3AN, UK8Géosciences, CNRS, Université Rennes, Campus de Beaulieu, 35042 Rennes, France9UMR6538, Institut Universitaire Européen de La Mer Place Nicolas Copernic 29280 Plouzané, France

ABSTRACT

We present seismic, core, log, and chrono-logic data on three early to middle Mio-cene sequences (m5.8, m5.4, and m5.2; ca. 20–14.6 Ma) sampled across a transect of seismic clinothems (prograding sigmoidal sequences) in topset, foreset, and bottomset locations beneath the New Jersey shallow con-tinental shelf (Integrated Ocean Drilling Pro-gram Expedition 313, Sites M27–M29). We recognize stratal surfaces and systems tracts by integrating seismic stratigraphy, litho-facies successions, gamma logs, and forami-niferal paleodepth trends. Our interpretations of systems tracts, particularly in the foresets where the sequences are thickest, allow us to test sequence stratigraphic models. Landward of the clinoform rollover, topsets consist of nearshore deposits above merged transgres-sive surfaces (TS) and sequence boundaries overlain by deepening- and fi ning-upward transgressive systems tracts (TST) and coars-ening- and shallowing-upward highstand sys-tems tracts (HST). Drilling through the fore-sets yields thin (<18 m thick) lowstand systems tracts (LST), thin (<26 m) TST, and thick HST (15–90 m). This contrasts with previously published seismic stratigraphic predictions of thick LST and thin to absent TST. Both HST and LST show regressive patterns in the cores. Falling stage systems tracts (FSST) are tenta-tively recognized by seismic downstepping, although it is possible that these are truncated

HST; in either case, these seismic geometries consist of uniform sands in the cores with a blocky gamma log pattern. Parasequence boundaries (fl ooding surfaces) are recognized in LST, TST, and HST. TS are recognized as an upsection change from coarsening- to fi ning-upward successions. We fi nd little evi-dence for correlative conformities; even in the foresets, where sequences are thickest, there is evidence of erosion and hiatuses asso-ciated with sequence boundaries. Sequence m5.8 appears to be a single million-year-scale sequence, but sequence m5.4 is a composite of 3 ~100-k.y.-scale sequences. Sequence m5.2 may also be a composite sequence, although our resolution is insuffi cient to demonstrate this. We do not resolve the issue of fractal ver-sus hierarchical order, but our data are con-sistent with arrangement into orders based on Milankovitch forcing on eccentricity (2.4 m.y., 405 and 100 k.y. cycles) and obliquity scales (1.2 m.y. and 41 k.y.).

INTRODUCTION

Sequence stratigraphy is based on recognition of unconformity-bounded sedimentary units on seismic profi les, in outcrop, in cored sections, and on geophysical logs (Vail et al., 1977; Van Wagoner et al., 1990). Sequences are objective units (e.g., Neal and Abreu, 2009), but the inter-pretation of sequences is often tied to genetic criteria (Mitchum et al., 1977), especially rela-tive sea-level change. The genetic connotation

remains controversial (e.g., Christie-Blick et al., 1988, 1990; Miall, 1991; Christie-Blick, 1991; Catuneanu, 2006; Embry, 2009). In addition, sequence nomenclature and approaches have proliferated, leading some to plead for a return to basics (Neal and Abreu, 2009). Basic principles of sequence stratigraphy focus on three stratal surfaces, i.e., sequence boundaries (SB), trans-gressive surfaces (TS), and maximum fl ooding surfaces (MFS), and stacking patterns of para-sequences (those bounded by fl ooding surfaces) and the attendant trends observed in cores as deepening- and shallowing-upward successions (Fig. 1). They are not explicitly tied to a rela-tive sea-level curve. We adopt a back to basics approach using new drilling data to address the architecture of seismic and core sequences.

A series of publications by Exxon Produc-tion Research Company illustrated sequences as sigmoidal, slug-shaped units with thin top-sets, thick foresets, and thin bottomset deposits bounded by sigmoidal clinoformal unconformi-ties and correlative conformities (Fig. 1; Vail, 1987; Van Wagoner et al., 1987; Posamentier and Vail, 1988; Posamentier et al., 1988). We apply the term clinothem to Miocene seismic sequences imaged beneath the New Jersey con-tinental shelf (Figs. 2 and 3). Clinothems are packages of sediment that prograde seaward and are bounded by surfaces (in this case sequence boundaries) with distinct sigmoidal (clinoform) geometry. The clinothem topsets were originally termed as the shelf and the rollover point as the shelf break (Vail et al., 1977). This has created

For permission to copy, contact [email protected]© 2013 Geological Society of America

Geosphere; October 2013; v. 9; no. 5; p. 1236–1256; doi:10.1130/GES00884.1; 13 fi gures; 7 supplemental fi gures.Received 14 November 2012 ♦ Revision received 1 July 2013 ♦ Accepted 8 August 2013 ♦ Published online 13 September 2013

Results of IODP Exp313: The History and Impact of Sea-level Change Offshore New Jersey themed issue

Page 2: Testing sequence stratigraphic models by drilling Miocene … · 2013-12-10 · Drilling foresets and testing sequence stratigraphic models Geosphere, October 2013 1237 confusion

Drilling foresets and testing sequence stratigraphic models

Geosphere, October 2013 1237

confusion because the modern continental shelf-slope break is typically in 120–200 m of water, averaging 135 m off New Jersey (Heezen et al., 1959). Two-dimensional backstripping of the New Jersey margin showed that the structur-ally controlled continental shelf-slope break occurred in 100–300 m of water from the Late Cretaceous to Miocene ~60 km seaward of Integrated Ocean Drilling Program Expedition 313 Site M29 (Steckler et al., 1999; Mountain et al., 2010) and that the rollover features (also called depositional shelf breaks, a term we avoid because it evokes the modern shelf break) asso-ciated with Miocene clinoforms are shallower, different features than the continental shelf-slope break.

HST

HST

TSTArrows point in fining direction

Rollover

TopsetCoastal Plain

Toe of SlopeDistal Toe

TS

Clinothem Model

Foreset

Bottomset

MFSSB

SB

LST

Figure 1. Clinothem model; arrows point in fi ning (deepening) direction. SB—sequence boundary (red lines); TS—transgressive surface (blue lines); MFS—maximum fl ooding surface (green lines); LST—lowstand systems tract (brown); TST—transgressive systems tracts (green); and HST—highstand systems tract (light pink). Rollover is equivalent to depositional shelf break of several authors.

+ ++

++++

+

+

++

Cape May

1072

902

903904

1071

905

906

1073

OceanView

72°W73°75° 74°76°77°

38°

39°

41°N

40°

Ew9009CH0698

Oc270

Seismic Profiles

Existing DrillsitesDSDPExplorationAMCORODP Leg 150, 150X903

ODP Leg 174A, 174AX1072

+

+

+ +

+

+

++

+

M27

M29M28

DelmarvaPeninsula

pre-Cretace

ous outcr

op

2000

m

1000

m

3000

m

200

m

New Jersey Sea-Level Transect

++

IslandBeach

Atlantic Ocean

SeaGirt

Fort Mott

MillvilleAtlantic City

Ancora

Bass River

Cenozoic

outcrop

Cretaceous o

utcrop

Medford

IODP Leg 313M27DoubleTrouble

WilsonLake

New Jersey

6010

Shelf

Rise

6020

6011

Cape May Zoo

6009

SlopeBethany Beach

Figure 2. Generalized bathymetric location map of the New Jersey–Mid Atlantic Margin sea-level transect showing three generations of multichannel seismic data (cruises R/V Ewing Ew9009, R/V Oceanus Oc270, R/V Cape Hatteras CH0698), onshore coreholes, and offshore coreholes drilled by AMCOR (Atlantic Margin Coring Project; Hathaway et al., 1979), the Ocean Drilling Program (ODP), the Integrated Ocean Drilling Program (IODP), and the Deep Sea Drilling Project (DSDP).

Page 3: Testing sequence stratigraphic models by drilling Miocene … · 2013-12-10 · Drilling foresets and testing sequence stratigraphic models Geosphere, October 2013 1237 confusion

Miller et al.

1238 Geosphere, October 2013

M27

M28

M29

Cum

ulat

ive

perc

ent

010

0

Cum

ulat

ive

perc

ent

010

0

Cum

ulat

ive

perc

ent

010

0

030

0cps

0 30

0cps

0 50

0cps

o1

o.5

m6

m5

.8

m5

.7

m5

.6

m5

.6

m5

.45

m5

.47

m5

.4

m5

.4

m5

.34

m5

.33

m5

.32

m5

.3m

5.2

m5

.2m5

m4

m4

.1

m4

.1

m4

.2

m4

.3

m4

.4

m4

.5

m4

.5

m3

m1

m5

.3

m5

0

200

400

600

800

1000

0 200

400

600

800

1000

Depth, meters

1000

090

0080

0070

0060

0050

0040

0030

0020

0010

00 c

dp

NW

SE

Exp

edit

ion

313

Oc2

70 5

290

10 k

m

Fig

ure

3. M

ulti

chan

nel s

eism

ic p

rofi l

e O

c270

529

(R/V

Oce

anus

) in

regi

on o

f Int

egra

ted

Oce

an D

rilli

ng P

rogr

am E

xped

itio

n 31

3 pr

ovid

ing

a lo

ng d

ip p

rofi l

e fr

om n

orth

wes

t (N

W)

to s

outh

east

(SE

). S

ites

M27

–M29

are

loc

ated

on

this

lin

e. A

xes

are

dept

h in

met

ers

obta

ined

us

ing

the

revi

sed

velo

city

-dep

th fu

ncti

on (M

ount

ain

and

Mon

teve

rde,

201

2) a

nd c

omm

on d

epth

poi

nts

(1 c

dp =

6.2

5 m

); a

ppro

xim

ate

scal

e (i

n km

) is

giv

en o

n lo

wer

rig

ht. M

ajor

sei

smic

seq

uenc

e bo

unda

ries

and

sel

ect

intr

aseq

uenc

e re

fl ect

ors

(o1,

o.5

) ar

e sh

own;

all

othe

rs a

re

seis

mic

seq

uenc

e bo

unda

ries

. Sup

erim

pose

d ar

e th

e co

arse

fra

ctio

n cu

mul

ativ

e pe

rcen

t lit

holo

gy a

nd d

ownh

ole

gam

ma

logs

(in

cou

nts

per

seco

nd, c

ps)

as d

escr

ibed

in M

iller

et

al. (

2013

).

Page 4: Testing sequence stratigraphic models by drilling Miocene … · 2013-12-10 · Drilling foresets and testing sequence stratigraphic models Geosphere, October 2013 1237 confusion

Drilling foresets and testing sequence stratigraphic models

Geosphere, October 2013 1239

Subdivision of sequences into systems tracts has been explicitly tied to relative sea-level changes (Vail, 1987; Van Wagoner et al., 1987; Posamentier and Vail, 1988; Posamentier et al., 1988; Coe, 2003; Catuneanu, 2006) and inter-pretation of systems tracts is often needlessly highly model dependent. Systems tracts were defi ned as linked depositional systems (Brown and Fisher, 1977) that are used to subdivide sequences into lowstand systems tracts (LST), transgressive systems tracts (TST), and high-stand systems tracts (HST; Vail, 1987; Van Wagoner et al., 1987; Posamentier and Vail, 1988; Posamentier et al., 1988). The falling stage systems tract (FSST) is a fourth systems tract (Plint and Nummedal, 2000), although its recognition can be controversial with respect to the location of the associated overlying sequence boundary (see summary in Coe, 2003). The LST, TST, and HST are separated by two distinct stratal surfaces: the transgres-sive surface (TS) and the maximum fl ooding surface (MFS). We summarize systems tracts as they apply to siliciclastic environments, focus-ing on these surfaces.

The fundamental surface in sequence stra-tigraphy is the sequence boundary and its rec-ognition is of primary importance. Seismic stratigraphic criteria for the sequence boundary include onlap, downlap, toplap, and erosional truncation (Mitchum et al., 1977). Criteria from core observations include irregular con-tacts, rip-up clasts, other evidence of rework-ing, intense bioturbation, major facies changes, stacking pattern changes (e.g., changes in coarsening versus fi ning upward; Fig. 1) and evidence for hiatuses (Van Wagoner et al., 1987; Miller et al., 2013). Geophysical log cri-teria include recognition of stacking patterns, particularly of parasequences (those bounded by fl ooding surfaces, FS; Van Wagoner et al., 1987, 1990), and the association of large uphole gamma-log increases with sequence boundaries, although these also occur at MFS. Sequence-bounding unconformities often lose seismic stratigraphic expression when traced basinward and the term correlative conformity has been included in the defi nition of sequence as a surface traced from the unconformity to one that has “…no physical evidence of erosion or non-deposition and no signifi cant hiatus…” (Mitchum et al., 1977, p. 206).

The TS generally separates the LST below from the TST above. Where no LST deposits are present (as is often the case on topsets; Fig. 1), or in seismic data where thin LST sediments are below seismic resolution, the TS merges with the sequence boundary. The TS marks a change from progradational to retro-gradational seismic stratigraphic successions

and a change in cores from coarsening-upward to fi ning-upward successions (Fig. 1) in shelf depositional environments (though these pat-terns may be complicated in the nearshore set-ting), and may appear as a shift from regressive sands below to fi ner grained muds above (Vail, 1987; Van Wagoner et al., 1987, 1988, 1990; Posamentier and Vail, 1988; Posamentier et al., 1988). The TS is diachronous and often linked to local erosion associated with marine ravinement as shoreface erosion cannibalizes former barrier island deposits (Demarest and Kraft, 1987).

The MFS separates the TST from the HST. The MFS is recognized in seismic sections as a downlap surface, an upsection change from retrogrational to progradational successions in seismic profi les and outcrops, and in cores as a change from fi ning-upward to coarsening-upward successions (Fig. 1) (Vail, 1987; Van Wagoner et al., 1987, 1988; Posamentier and Vail, 1988; Posamentier et al., 1988). In cores, sediments deposited along the MFS usually record the deepest water of a sequence; fur-thermore, these sediments are often associ-ated with a condensed section recognized by intense bioturbation, in situ glauconite, phos-phorite, abundant organic carbon, greater mud versus sand, planktonic microfossils, and in situ shells (Loutit et al., 1988; Kidwell, 1989, 1991). The TST is transgressive (generally fi n-ing upsection; Fig. 1) and thus is associated with retrogradational parasequence sets, gener-ally stepping up onto the topsets of the previ-ous sequence (Fig. 1). The HST is regressive, associated with aggradational to progradational and degradational parasequence sets (Neal and Abreu, 2009), downlaps on the MFS, and is generally overlain by the upper sequence boundary (Vail, 1987; Van Wagoner et al., 1987, 1988; Posamentier and Vail, 1988; Posa-mentier et al., 1988).

Interpretation of the LST is controversial because of the uncertainties in placement of its base versus the FSST (Coe, 2003), the var-ied facies it contains, and the fact that it is the one salient feature separating sequences from transgressive-regressive cycles (Christie-Blick and Driscoll, 1995; Catuneanu et al., 2009; Embry, 2009). Vail et al. (1977) fi rst termed all strata that onlap seaward of the clinothem rollover (Fig. 1; his shelf break) as lowstand deposits. Subsequent studies have defi ned the LST in terms of sea-level curves (Vail, 1987; Van Wagoner et al., 1987, 1988; Posamentier and Vail, 1988; Posamentier et al., 1988; Coe, 2003), engendering debate. There is general agreement that sediments of the LST directly overlie the sequence boundary, are the lower regressive systems tract containing progra-

dational to aggradational parasequence sets, and generally coarsen up to the TS (Vail, 1987; Van Wagoner et al., 1987; Posamentier et al., 1988; Coe, 2003; Neal and Abreu, 2009). How-ever, there has been a tendency to attribute all coarse-grained sediments overlying a sequence boundary to those of the LST, even when unjus-tifi ed (e.g., transgressive estuarine gravels inter-preted as lowstand deposits; Christie-Blick and Driscoll, 1995).

In the FSST, strata not only prograde as they do in the underlying HST, they also step down into the basin (often with sharp-based sands) and offl ap progressively seaward (Plint and Nummedal, 2000), with progradation and progressively steepening foresets (e.g., Proust et al., 2001). The FSST is partially equivalent to the forced regression of Posamentier et al. (1992) and contrasts with the HST, where strata progressively onlap landward (Plint and Nummedal, 2000). A distinct surface separat-ing the FSST from the underlying HST may be lacking (Plint and Nummedal, 2000). How-ever, in many cases there is a marine erosion surface–associated regression (Proust et al., 2001), especially associated with Pleistocene 100 k.y. sequences (e.g., Trincardi and Correg-giari, 2000; Rabineau et al., 2005). In general, the sequence boundary is placed at the top of the FSST (Plint and Nummedal, 2000), although “...there is still some controversy as to where the sequence boundary should be placed” (Coe, 2003, p. 86).

Most sequence stratigraphic interpretations rely heavily on links to hypothetical relative sea-level curves (see summary in Catuneanu et al., 2009). Early models interpreted deposition of (1) the LST from the time of maximum rate of relative and/or eustatic fall (falling infl ection point) associated with the sequence boundary to the beginning of the rise (Posamentier et al., 1988); (2) the TST from the beginning of the rise to about the time of the maximum rate of rise at the MFS (Galloway, 1989); and (3) the HST from the maximum rate of rise to the time of maximum rate of fall (Posamentier and Vail, 1988). Subsequent publications have devel-oped strikingly different timings (i.e., with the LST lagging a quarter cycle and starting at the beginning of the rise, MFS late in the relative rise) of systems tracts relative to hypothetical sea-level curves (e.g., Coe, 2003; Catuneanu et al., 2009; http://www.sepmstrata.org/page.aspx?&pageid=32&3). However, application of any model is an oversimplifi cation because position of a stratal surface relative to a sea-level curve is a function of preexisting geom-etry, rates of subsidence (including differen-tial subsidence that precludes computation of a single relative sea-level curve), and sedi-

Page 5: Testing sequence stratigraphic models by drilling Miocene … · 2013-12-10 · Drilling foresets and testing sequence stratigraphic models Geosphere, October 2013 1237 confusion

Miller et al.

1240 Geosphere, October 2013

ment supply (including shifting depocenters) ( Christie-Blick et al., 1990).

The controversial nature of the LST ( Christie-Blick, 1991; Christie-Blick and Driscoll, 1995) and the FSST (see summary in Coe, 2003) have led some to return to interpreting sequences as largely transgressive-regressive (T-R) cycles (Embry, 2009). T-R cycles describe sequences where lowstand deposits are absent, including many outcrop sections. For example, T-R cycles typify onshore New Jersey coastal plain depo-sition (e.g., Owens and Gohn, 1985), where the TS and sequence boundary are generally merged (Olsson et al., 1987; Sugarman et al., 1993; Miller et al., 1998). Similar T-R cycles have been interpreted in Europe (e.g., Hancock, 1993) and the western interior of the U.S. (e.g., Hancock and Kauffman, 1979). However, thin (<1 m) regressive LST can be preserved even on clinothem topsets of the New Jersey coastal plain (Miller et al., 1998; Browning et al., 2008), and geometries of forced regression, FSST, and lowstand deposits must be considered on the clinothem foresets. On the foresets, it is not an option to rely solely on T-R cycles, because low-stand deposits occur above sequence boundaries (Fig. 1).

Neal and Abreu (2009) focused on the basic stratal surfaces (SB, TS, and MFS) and stack-ing patterns of parasequence sets, following Mitchum and Van Wagoner (1991) in noting that sequences are scale independent. They identifi ed systems tracts by distinguishing the following stacking patterns in cores and outcrop. (1) LST are progradational to aggradational (coarsening upward, ending in largely structureless sand; Fig. 1). (2) TST are retrogradational (fi ning upward; Fig. 1). (3) HST are aggradational to progradational and degradational (coarsening upward). Neal and Abreu (2009) noted that LST may be found landward of the rollover (deposi-tional shelf edge). We adopt their approach of focusing on SB, TS, MFS and stacking and/or water depth trends using seismic-core-well log integration offshore of New Jersey.

The New Jersey margin has several genera-tions of multichannel seismic data (MCS) that have imaged clinothem sequences (fi rst called prograding deltas; Schlee, 1981). Greenlee et al. (1988) and Greenlee and Moore (1988) used industry seismic profi les to showcase the New Jersey shelf as a classic example of Miocene prograding sequences. Greenlee et al. (1992) interpreted the presence of thick lowstand wedges and HST, seismically lack-ing TST, for Miocene sequences beneath the middle to outer continental shelf of New Jer-sey. Poulsen et al. (1998) investigated middle Miocene sequences imaged in higher resolu-tion seismic across the New Jersey outer con-

tinental shelf and reached a similar interpreta-tion of only LST and HST. Monte verde et al. (2008) and Monte verde (2008) focused on Mio-cene sequences discussed here (ca. 23–13 Ma) that are landward of the middle to outer shelf seismic profi les of Greenlee et al. (1988) and Poulsen et al. (1998), and similarly concluded that sequences were almost approximately equal thicknesses of LST and HST, and that TST was either below seismic resolution or completely absent. The early to early-middle Miocene seis-mic sequences (discussed in Monte verde et al., 2008; Monteverde, 2008) were sampled by Inte-grated Ocean Drilling Program (IODP) Expedi-tion 313 (Figs. 2 and 3; Supplemental Fig. 11), with continuous cores and geophysical logs.

IODP Expedition 313 was designed to test sequence stratigraphic relationships across a series of early to middle Miocene clinothems (Figs. 2 and 3; Mountain et al., 2010); 15 early to middle Miocene (ca. 23–13 Ma) seismic sequence boundaries were recognized using criteria of onlap, downlap, erosional truncation, and toplap (Monteverde et al., 2008; Monte-verde, 2008; Mountain et al., 2010). Core recovery was very good (~80%) considering the challenges in coring shallow-water sands and geophysical logs were obtained at all three sites. Sequence boundaries in cores and logs were recognized based on integrated study of key core surfaces, lithostratigraphy and process sedimentology (grain size, mineralogy, facies, and paleoenvironments), facies successions, benthic foraminiferal water depths, downhole logs, core gamma logs, and chronostratigraphic ages (Mountain et al., 2010; Miller et al., 2013). Velocity and density logs allow construction of synthetic seismograms at Sites M27 and M29 (Mountain and Monteverde, 2012), providing fi rm placement of sequence boundaries (Miller et al., 2013) and a starting point for deciphering systems tracts. Ages of sequences and hiatuses are derived by integrating Sr isotope stratigra-phy and biostratigraphy (diatoms, nanno fossils, and dinocysts) on age-depth diagrams with a resolution of ±0.25 to ±0.5 m.y. (Browning et al., 2013). In this contribution we focus on three sequences sampled across of full range of topset, foreset, and bottomsets: sequences m5.8, m5.4, and m5.2.

The objective of this paper is to integrate seismic interpretations done before drilling

(Greenlee et al., 1988, 1992; Monteverde et al., 2008; Monteverde, 2008) with those done subsequently (Mountain et al., 2010; this study) and with core and geophysical log data to provide new insights into the interpretations of systems tracts focusing on critical thick foreset deposits (Figs. 4–11). We recognize stratal surfaces and systems tracts by integrat-ing seismic stratigraphic interpretation, litho-facies successions, gamma logs, and benthic foraminiferal paleodepth trends. Our inter-pretation of systems tracts across the three clinothems allows us to test sequence strati-graphic models.

METHODS

Seismic Interpretation

Seismic sequence boundaries m5.8, m5.4, and m5.2 were identified in multichannel seismic grids obtained on R/V Ewing cruise Ew9009, R/V Oceanus cruise Oc270, and R/V Cape Hatteras cruise CH0698 (in 1990, 1995, and 1998, respectively; Monteverde et al., 2008; Monteverde, 2008; Mountain et al., 2010). We focus here on interpretations of Oc270 line 529, which crosses Sites M27, M28, and M29 (Figs. 2 and 3; Supplemental Fig. 1 [see footnote 1]). Seismic sequence boundaries m5.8, m5.4, and m5.2 were identifi ed based on refl ector termina-tions (onlap, downlap, erosional truncation, and toplap) on multiple lines and loop correlated throughout the seismic grids (Fig. 2). These criteria allow differentiation of these sequence boundaries from surfaces associated with FSST or truncated HST (e.g., refl ectors 2 and 3 in Fig. 7). Sequences are named according to their basal refl ector boundary, such that refl ec-tor m5.8 is the base of sequence m5.8. Several additional refl ectors that are potential sequence boundaries (m5.34, m5.33, and m5.32; Fig. 3) were identifi ed within sequence m5.4 (Fig. 3) by two of us (D. Monteverde and G. Mountain, in Mountain et al., 2010), but not loop correlated; their stratal signifi cance is discussed herein. We trace internal refl ectors within sequences m5.8, m5.4, and m5.2. MFS (green lines, Figs. 4–11) are seismically recognized by signifi cant down-lap across the sequence and onlap near to or landward of the rollover (Fig. 1); in sequences where there is more than one downlap surface, the strati graphically lowest is taken as the seis-mic MFS. Seismic criteria alone are insuffi cient to unequivocally recognize TS, and placement of TS was done by iteration with core studies (see following). In all three cases, TS (blue lines, Figs. 4–11) onlap the basal sequence bound-ary seaward of the rollover and farther seaward downlap onto the sequence boundary or merge

1Supplemental Figure 1. (A) Uninterpreted MCS profi le Oc270 Line 529 sized to print at 18 × 36 inches. (B) Interpreted MCS profi le Oc270 Line 529 sized to print at 18 × 36 inches. If you are viewing the PDF of this paper or reading it offl ine, please visit http://dx.doi.org/10.1130/GES00884.S1 or the full-text article on www.gsapubs.org to view Supple-mental Figure 1.

Page 6: Testing sequence stratigraphic models by drilling Miocene … · 2013-12-10 · Drilling foresets and testing sequence stratigraphic models Geosphere, October 2013 1237 confusion

Drilling foresets and testing sequence stratigraphic models

Geosphere, October 2013 1241

NG

R (

cps)

M28

147R

148R

149R

150R

151R

152R

153R

154R

155R

156R

157R

158R

159R

160R

161R

162R

163R

164R

165R

166R

167R

168R

169R

170R

171R

670

660

650

640

630

620

610

600

gg

gg

wd

wd

gw

d

wd

p

p

p wd

xx

xwd

p

xx

wd

wd

xxxx

xxxx

x

xxxx

xxx

Lith

olog

yC

ore

reco

very

Depth (mcd)

River-influencedoffshore

Toe slope

Rive

r-in

fluen

ced

offs

hore

50 5025

Pal

eode

pth

IntegratedWaterdepth

75

5010

00

Cum

ulat

ive

Lith

olog

y (%

)N

GR

(cp

s)

05

1015

2025

TG

R (

cps)

010

020

030

0

m5.

761

1.6

m5.

866

2.98

18.8

19.5

20.0

?20.

5

M27

TG

R (

cps)

Lith

olog

yC

ore

reco

very

gg

xxxx

xxxx

xxxx

xxx

xxxxxxxxxxxxxxx

XX

X

X

X

xxxx

xxxx

xxxx

xxx

XXXX

XXXX

XX

XXXX

XXXX

XXXX

XXXX

XXXX

g

g

Depth (mcd)

73

0

74

0

75

0

76

0

207R

208R

209R

210R

211R

212R

213R

214R

215R

216R

217R

Dysoxic prodelta

OFF

75-1

00 m

50-8

0 m

50-8

0 m

75-1

00 m

50-1

00 m

Environment

Pal

eode

pth

BenthicBiofacies

IntegratedWaterdepth

Cum

ulat

ive

Lith

olog

y (%

)

50

004

812

Reflector

Depth (mcd)

Age (Ma)

m5.

7

m5.

8

728.

56

746/

753.

80

7575 75

18.8

20.0

20.2

20.5

100

0

M29

Pal

eode

pth

SF River-influenced SOT River-influenced OFF

50-8

0 m

50-8

0 m

50-6

0 m

50-8

0 m

50-8

0 m

50-8

0 m

50-8

0 m

50-8

0 m

50-6

0 m

50-6

0 m

50-6

0 m

50-8

0 m

75-8

0 m

75-8

0 m

75-8

0 m

50-8

0 m

Toe-of-slopeapronEnvironment

BenthicBiofacies

IntegratedWaterdepth

5 10 30 75 75 6075 35 50

HS

T

LST

TS

T

Systems

TG

R (

cps)

NG

R (

cps)

020

040

0

05

1015

2025

Reflector

Depth (mcd)

Age (Ma)

171R

172R

173R

174R

175R

XX

XX

XX

XX

XX

XX

XX

XX

XX

XX

XX

XX

XX

XX

XX

XX

XX

XX

XX

XX

XX

XX

XX

XX

XX

XX

XX

XX

XX

XX

XX

XX

XX

XX

XX

XX

XX

XX

XX

XX

XX

XX

XX

XX

XX

XX

XX

XX

XX

XX

XX

XX

XX

XX

XX

XX

XX

XX

XX

XX

XX

XX

XX

XX

XX

XX

XX

XX

XX

XX

XX

XX

XX

XX

XX

XX

XX

gg g

gg

g

500

127R

128R

129R

130R

131R

132R

133R

134R

135R

136R

137R

138R

139R

140R

141R

142R

143R

144R

145R

146R

147R

148R

149R

150R

151R

152R

153R

154R

155R

156R

157R

158R

159R

160R

161R

162R

163R

164R

165R

166R

167R

168R

169R

170R

XX

XX

XX

XX

XX

XX

wd g

gg

wd

wd

wd w

d

gg

ggg

gg

gg

gLith

olog

yC

ore

reco

very

Depth (mcd)360

370

380

390

400

410

420

430

440

450

460

470

480

490

5010

00

Cum

ulat

ive

Lith

olog

y (%

)

19.2

20.1

20.7

m5.

7/“9

”36

1.28

m5.

8/“1

”49

4.87

?MF

S

?MF

S

?MF

S

TS

Seq

uenc

e m

5.8

For

eset

Pro

delta

Bot

tom

set

Pro

delta

Bot

tom

set

477.

52

457.

78

451.

36

442

Sei

smic

Lith

olog

y

Ben

thic

MaximumFlooding Zone

Evironment

654

742

?MF

S?M

FS

Reflector

Depth (mcd)

Age (Ma)

“5”

“3”

TS

T/

LST

HS

T

TS

T

HS

T

75-1

00 m

cs

gvf

vcS

and

cs

gvf

vcS

and

cg

vfvc

San

d

Gla

ucon

ite

Cla

y/si

lt

Ver

y fin

e an

d fin

e qu

artz

For

amin

ifers

/she

lls

Cum

ulat

ive

Lith

olog

y

Mic

a

Oth

er

Med

ium

& c

oars

er q

uart

z

XX

Indu

rate

d/no

dule

Cro

ss la

min

ae

Wed

ge-p

lana

r la

min

ae

Bio

turb

atio

n

Con

volu

ted

bedd

ing

Sco

ur

Hor

izon

tal b

eddi

ng

gG

lauc

onite

pP

yrite

Sw

aley

cro

ss s

trat

ifica

tion

Hum

moc

ky c

ross

str

atifi

catio

n

Org

anic

mat

ter

wd

She

ll fr

agm

ents

Art

icul

ated

she

ll

For

amin

ifers

Gas

trop

ods

She

ll de

bris

Bur

row

Ech

inoi

ds

Sym

bols

Cla

y

Silt

Very

fine

/fine

san

d

Coa

rse/

med

ium

san

d

Gla

ucon

ite (

>25

%)

Bio

clas

tic h

oriz

on

Lith

olog

yK

EY X

XIn

dura

ted/

nodu

le

400

200

Fig

ure

4. C

ompa

riso

n of

seq

uenc

e m

5.8

at I

nteg

rate

d O

cean

Dri

lling

Pro

gram

Exp

edit

ion

313

Site

s M

27, M

28, a

nd M

29, s

how

ing

core

de

pths

in

met

ers

com

posi

te d

epth

(m

cd),

cor

e nu

mbe

r (1

H–2

1H,

whe

re H

ind

icat

es r

ecov

ery

by h

ydra

ulic

pis

ton

cori

ng;

R—

rota

ry;

and

X—

exte

nded

cor

e ba

rrel

); c

ore

reco

very

(gr

ay—

reco

vere

d, w

hite

—ga

p);

litho

logy

(c—

clay

; s—

silt

; vf

—ve

ry fi

ne

sand

; vc

—ve

ry

coar

se s

and;

g—

grav

el a

nd/o

r pe

bble

s);

coar

se f

ract

ion

cum

ulat

ive

perc

ent

litho

logy

(br

own—

mud

; lig

ht y

ello

w—

fi ne

quar

tz s

and;

dar

k ye

llow

—m

ediu

m-c

oars

e qu

artz

san

d; g

reen

—gl

auco

nite

san

d; b

lue—

carb

onat

e [s

hells

and

for

amin

ifer

a];

arro

ws

poin

t in

fi n

ing

dire

c-ti

on);

sym

bols

on

key

at u

pper

rig

ht (

afte

r M

ount

ain

et a

l., 2

010)

. Env

iron

men

tal

inte

rpre

tati

on b

ased

on

litho

faci

es i

s af

ter

Mou

ntai

n et

al.

(201

0) (

SF—

shor

efac

e; S

OT

—sh

oref

ace-

offs

hore

tra

nsit

ion

[low

er s

hore

face

]; O

FF

—of

fsho

re).

Pal

eow

ater

dep

ths

(in

met

ers)

are

ba

sed

on b

enth

ic f

oram

inif

eral

bio

faci

es a

fter

Kat

z et

al.

(201

3). I

nteg

rate

d pa

leow

ater

dep

ths

(in

met

ers)

are

bas

ed o

n be

nthi

c fo

ram

i-ni

fera

l bio

faci

es a

nd li

thof

acie

s af

ter

Mill

er e

t al

. (20

13).

Gam

ma

logs

: re

d—do

wnh

ole

wir

elin

e m

easu

rem

ents

as

tota

l gam

ma

ray

(TG

R)

in c

ount

s pe

r se

cond

(cp

s);

blue

dot

s—na

tura

l ga

mm

a-ra

y (N

GR

) m

easu

rem

ents

mad

e on

uns

plit

who

le c

ores

, sc

ale

in c

ps;

data

aft

er

Mou

ntai

n et

al.

(201

0). R

efl e

ctor

s: r

ed—

sequ

ence

bou

ndar

y; b

lue—

tran

sgre

ssiv

e su

rfac

e (T

S);

gree

n—m

axim

um fl

oodi

ng s

urfa

ce (

MF

S).

Das

hed

red

lines

—un

cert

ain

plac

emen

t of

seq

uenc

e bo

unda

ry. L

ST—

low

stan

d sy

stem

s tr

act;

TST

—tr

ansg

ress

ive

syst

ems

trac

t; H

ST—

high

stan

d sy

stem

s tr

act;

FSS

T—

falli

ng s

tage

sys

tem

s tr

act.

Age

s fo

r su

rfac

e im

med

iate

ly b

elow

and

abo

ve s

eque

nce

boun

dari

es a

re a

fter

B

row

ning

et

al. (

2013

).

Page 7: Testing sequence stratigraphic models by drilling Miocene … · 2013-12-10 · Drilling foresets and testing sequence stratigraphic models Geosphere, October 2013 1237 confusion

Miller et al.

1242 Geosphere, October 2013

with the MFS. Other internal refl ections were traced (yellow lines, Figs. 5, 7, and 10) and used to interpret stacking patterns and to construct age-distance plots (top panels of Figs. 5, 7, and 10; also called Wheeler diagrams or chronostrati-graphic charts of Vail et al., 1977).

Sequences, Lithology, and Paleoenvironments in Cores and Core-Seismic Integration

Sequence boundaries in the Expedition 313 cores were recognized on the basis of physical stratigraphy and age breaks (Mountain et al., 2010; Miller et al., 2013). Criteria for recog-nizing sequence-bounding unconformities in

coreholes (e.g., Browning et al., 2006) that were applied to Expedition 313 cores include: (1) irregular contacts, with as much as 5 cm of relief on a 6.2-cm-diameter core; (2) rework-ing, including rip-up clasts found above the contact; (3) intense bioturbation, including bur-rows fi lled with overlying material; (4) major litho facies shifts and changes in stacking pat-tern (discussed in the following); (5) upsection gamma-ray increases associated with changes from low-radioactivity sands below to hotter clays or glauconite sands immediately above sequence boundaries, and/or marine omission surfaces (e.g., with high U/Th scavenging); (6) shell lags above the contact; and (7) age breaks indicated by Sr isotope stratigraphy or

biostratigraphy. Numerous sequence boundaries are illustrated in core photographs (Miller et al., 2013). A velocity versus depth function was used to make initial seismic-core correlations of seis-mic sequence boundaries to core surfaces iden-tifi ed from visual evidence (core descriptions and photographs) and log data (Mountain et al., 2010; Mountain and Monteverde, 2012; Miller et al., 2013). Synthetic seismo grams from Sites M27A and M29A (Mountain and Monteverde, 2012) provide a check on seismic-core correla-tions and predicted depths of seismic sequence boundaries. The resultant seismic-core-log cor-relations (summarized in Miller et al., 2013) were used to construct site to site correlations for the three sequences m5.8, m5.4, and m5.2 that

(m6) 1(m5.8) 34

56

78

9(m5.7)

0

(m5.8) 1

345678

(m5.7) 9

TS

MFS

HST

TST

LST

20.720.1

19.2?

Age(Ma)

SystemsTract

Onlap

Onlap

DownlapDownlap

Downlap

19.7

0.500

0.600

TWTT

(sec

)

M27

2

–1

–2–3

CDP 8000 7000

2

0–1–2–3

1(m5.8)

3456789(m5.7)

2

0–1–2–3

SB

Reflector

Sand and Sandy mud

Mud and muddy sand

Mud

0 5 km

SB

Figure 5. Interpreted seismic profi le and Wheeler diagram (stratigraphic position versus distance; Wheeler, 1958) of sequence m5.8 across the foreset at Integrated Ocean Drilling Program Expedition 313 Site M27. Bottom panel is interpreted seismic profi le in two-way travel-time (TWTT, in seconds versus cdp, common depth point). LST—lowstand systems tract; TST—transgressive systems tract; HST—high-stand systems tract; FSST—falling stage systems tract; MFS—maximum fl ooding surface; SB—sequence boundary. Red arrows indicate refl ector terminations; refl ectors in red indicate sequence boundaries; refl ectors in blue indicate TS; and refl ectors in green indicate MFS. Other internal refl ections are indicated in shades of yellow. Cumulative lithology is superimposed on the site. Arbitrary numbers assigned to refl ectors are used to construct a time-distance plot at the top; scale of the Wheeler diagram on left assumes constant ages between refl ec-tors; age estimates (shown in Ma) are derived from Browning et al. (2013).

Page 8: Testing sequence stratigraphic models by drilling Miocene … · 2013-12-10 · Drilling foresets and testing sequence stratigraphic models Geosphere, October 2013 1237 confusion

Drilling foresets and testing sequence stratigraphic models

Geosphere, October 2013 1243

sampled topsets, foresets, and bottomsets in the three coreholes (Figs. 4, 6, and 9).

Lithologic trends are essential in interpreting systems tracts. The Expedition 313 sedimen-tologists produced visual core descriptions and differentiated clay, silt, and various sand frac-tions visually and using smear slides (Moun-tain et al., 2010). These lithologic descriptions have been synthesized into general lithology columns (essentially unchanged from Moun-tain et al., 2010) and presented as lithology in Figures 4, 6, and 8. Quantitative and qualitative lithology data were added (Miller et al., 2013) and weight percent mud (<63 μm), very fi ne and fi ne sand (63–250 μm), and medium sand and coarser sediment (>250 μm) were mea-sured in washed samples at ~1.5 m intervals; the abundance of glauconite, shells, and mica in the sand fraction (>63 μm) was semiquan-titatively determined by splitting 1727 samples into aliquots and visually estimating percent-ages on a picking tray. The data (presented as cumulative lithology in Figs. 4, 6, and 9) clearly show distinct trends in grain size and mineral-ogy that complement and extend the lithology columns presented as visual core descriptions (in Mountain et al., 2010).

Paleoenvironments are interpreted from litho-facies and biofacies. Lithofacies successions are interpreted using a wave-dominated shoreline model (summarized in Mountain et al., 2010), recognizing upper shoreface (0–5 m water depth), lower shoreface (5–10 m), shoreface-offshore transition (10–20 m), and offshore (>30 m) environments. Other environmental information (e.g., river-dominated) are from Mountain et al. (2010). Benthic forami niferal biofacies were reported in Mountain et al. (2010) and in greater detail in Katz et al. (2013). Ben-thic foraminifera provide paleodepth constraints following the general paleo bathy metric model of Miller et al. (1997) for coeval onshore New Jersey sections. In general, innermost neritic (<10 m) sediments were barren or yielded only Lenticulina spp., Hanzawaia hughesi–dominated biofacies are 10–25 m, Nonionella pizarrensis–dominated biofacies are 25–50 m, Bulimina gracilis–domi nated bio facies are 50–80 m, Uvigerina spp.-dominated biofacies are 75–100 m, and high-diversity, low-domi-nance assemblages with key indicator taxa (e.g., Cibici doides pachyderma, Hanzawaia man-taensis, and Oridor salis) are >100 m (Mountain et al., 2010; Katz et al., 2013). In addition, plank-tonic foraminiferal abundance changes provide an additional proxy for water-depth variations at the Expedition 313 sites, with increasing percentages of planktonic forami nifera of total foraminifera with increasing water depth (Katz et al., 2013). We present both benthic forami-

100

0

175R

176R

177R

178R

179R

180R

181R

182R

183R

184R

66

0

65

0

64

0xx

xxxx

xxx

XX

XX

XX

XX

XX

wd g

XX

XX

XX

XX

XX

XX

XX

X

wd

gg

xxxxxxxxxxxxxxx

gg

g

xxxx

xxxx

x

Lith

olog

y

Bas

ed o

n ag

es

Bas

ed o

n ag

es

Cor

ere

cove

ry

Toe

ToeEnvironment

Pal

eode

pth

BenthicBiofacies

IntegratedWaterdepth

70-8

0 m

75-1

00 m

75-1

00 m

75-1

00 m

75-1

00 m

Cum

ulat

ive

Lith

olog

y (%

)

50

0

TG

R (

cps)

NG

R (

cps)

04

812

Reflector

Depth (mcd)

Age (Ma)

643.

19

662.

37

70 75 75 75

16.1

17.6

17.7

17.7

200

400

490

480

470

460

450

440

430

420

410

400

390

380

370

360

56R

57R

58R

59R

60R

61R

62R

63R

64R

65R

66R

67R

68R

69R

70R

71R

72R

73R

74R

76R

77R

78R

79R

80R

81R

82R

83R

84R

85R

86R

87R

88R

89R

90R

91R

92R

93R

94R

95R

96R

97R

98R

99R

100R

101R

102R

103R

104R

105R

g

g

gg

g g

g

gg

g

g

g xxxx

xxx

gwd

? ? ? C.D.

C.D.

wd wd

wd

wd

wd

wd

x

wd

wd

wd wd

wd wd

wd

Depth (mcd)

500

106R

wd

wd

107R

108R

109R

110R

111R

112R

520

510

wd

wd

wd gLith

olog

yC

ore

reco

very

25-5

0 m

50-6

0 m

50-7

5 m

50-7

5 m

75-1

00 m

75-1

00 m

50-7

5 m

SF River-influencedSF

River-influencedSOT

River-influencedSOT

River-influencedOFF

OFF

10-2

5 m

10-2

5 m

10-2

5 m

10-2

5 m

10-2

5 m

10-2

5 m

10-2

5 m

15 10 1025 25 30 50 50 50 75 50 5075

Toe slopeapron

50-1

00 m

10-2

5 m

Pal

eode

pth

BenthicBiofacies

IntegratedWaterdepth

Systemstracts

Environment

Cum

ulat

ive

Lith

olog

y (%

)50

100

0

TG

R (

cps)

010

020

030

0

NG

R (

cps)

04

812

m5.

4/m

5.4-

1“4

m5.

34/

“7”

m5.

33/

“12”

361

405

449

479

512.

33

16.6

17.7

17.6

17.4

16.7

17.6

17.9

16.3

Reflector

Depth (mcd)

Age (Ma)

391

“8”

m5.

32/

“14”

m5.

3/“2

1”

Seq

uenc

e m

5.4

For

eset

Top

set

75-1

00 m

Bot

tom

set

M27

M28

M29

HS

T

HS

T

HS

T

LST

TS

T

TS

T

LST

LST T

S

475

m5.

35“6

”49

4

501

“5”

MF

S

MF

S

TS

MF

S

“10”

TS

393

m5.

4/m

5.4-

1

Depth (mcd)

87R

88R

89R

90R

91R

92R

93R

94R

95R

96R

97R

98R

99R

100R

Depth (mcd)

270

250

260

280

290

101R

102R

103R

104R

gg

**

*

300

Cor

ere

cove

ryLi

thol

ogy

75 m

25 m

75 m

25-5

0 m

25-5

0 m

25-5

0 m

SF

SO

T

OT

OF

F

OF

FEnvironment

Pal

eode

pth

BenthicBiofacies

IntegratedWaterdepthSystemsTract

35 75 30 403045 15 40 50

HS

T

HS

T

TS

T40

-50

m25

m

S

5010

00

Cum

ulat

ive

Lith

olog

y (%

)

m5.

3

m5.

3

249.

76

256.

19

295.

01

15.8

16.5

17.0

17.7

15.8

16.5

16.6

16.9

Reflector

Depth (mcd)

Age (Ma)

m5.

3327

1.23

m5.

34/

m5.

4m

erge

d

Pre

ferr

ed

TG

R (

cps)

020

040

060

0

NG

R (

cps)

05

1015

2025

150

MF

S

MF

S

?TS

TS

T?

35-5

0 m

m5.

3/m

5.34

mer

ged

288

265

cs

gvf

vcS

and

cs

gvf

vcS

and

cs

gvf

vcS

and

Fig

ure

6. C

ompa

riso

n of

seq

uenc

e m

5.4

at I

nteg

rate

d O

cean

Dri

lling

Pro

gram

Exp

edit

ion

313

Site

s M

27, M

28, a

nd M

29. C

apti

on a

nd k

ey a

s in

Fig

ure

4.

Page 9: Testing sequence stratigraphic models by drilling Miocene … · 2013-12-10 · Drilling foresets and testing sequence stratigraphic models Geosphere, October 2013 1237 confusion

Miller et al.

1244 Geosphere, October 2013

21

3

1516

10

0

21(m

5.3)

89

12(m

5.33

)13

14(m

5.32

)20

1

(m5.

4) 4 3

5 (m

5.37

)689101113

(m5.

32) 1

415161718 01920 214

(m5.

4)367

(m5.

34)

89101112 (m

5.33

)1314

(m5.

32)

15161718 01920 221 (m

5.3)

Topl

ap

FSS

T?

Onl

ap

??E

rosi

on

Onl

ap

SB

?O

nlap

Topl

ap/e

rosi

on

Ero

sion

Ero

sion

Dow

nlap

Ero

sion

Ero

sion

Ero

sion

Dow

nlap

Dow

nlap

SB SB

(m5.

34) 7

(m5.

33) 1

2

(m5.

3) 2

1

17.7

Ma

17.6

Ma

17.6

Ma

17.4

Ma

16.7

Ma

16.6

Ma

16.3

Ma

17.9

Ma

Ero

sion

4(m

5.4)5

6

7(m

5.34

)

10

1718

19

11

7000

6000

5000

CD

P

TSSBTSMFS

MFS

TST

MFS

HS

T

LST

LST

TST

HS

T

TST

HS

T

FSS

T?

Age

(Ma)

Sys

tem

sTr

act

0.40

0

0.50

0

0.60

0

TWTT (sec)

M27

M28

m5.

45

Mer

ged

sequ

ence

bou

ndar

y an

d M

FS

5

Ref

lect

or

San

d an

d sa

ndy

mud

Mud

and

mud

dy s

and

Mud

San

d (in

ferr

ed)

05

km

Fig

ure

7. I

nter

pret

ed s

eism

ic p

rofi l

e an

d W

heel

er d

iagr

am o

f seq

uenc

e m

5.4

acro

ss th

e fo

rese

t at I

nteg

rate

d O

cean

Dri

lling

Pro

gram

Exp

e-di

tion

313

Sit

e M

28, e

xten

ding

to

the

tops

et a

t Si

te M

27. C

apti

on a

s in

Fig

ure

5.

Page 10: Testing sequence stratigraphic models by drilling Miocene … · 2013-12-10 · Drilling foresets and testing sequence stratigraphic models Geosphere, October 2013 1237 confusion

Drilling foresets and testing sequence stratigraphic models

Geosphere, October 2013 1245

niferal paleodepths and integrated paleodepths obtained by combining lithofacies and bio facies constraints (Figs. 4, 6, and 9) and percent plank-tonic foraminiferal data.

We present gamma-log values obtained downhole through the drill pipe (total gamma ray, TGR) and those obtained directly on the

core in the laboratory (natural gamma ray, NGR) (Figs. 4, 6, and 9). Gamma-log data record litho-logic variations primarily of quartz sands versus clays or glauconite-rich sedi ments, with low gamma readings in sands and high gamma-log values in muds, and generally highest values in glauconite-rich sediments.

Here we interpret TS, MFS, and systems tracts in sequences identified by Mountain et al. (2010) and updated in Miller et al. (2013, including detailed justifi cation of placement of sequence boundaries). In cores, MFS are rec-ognized by an uphole change in pattern from deepening-upward (generally fi ning upward)

0 20 40 60 80 100

420

430

440

450

460

470

Dep

th (

mcd

)

81R

82R

83R

84R

85R

86R

87R

89R

90R

91R

92R

93R

95R

96R

94R

88R

wd

wd

wd

wd

x

wd

Paleodepth50-60 m @ 430.99 mcd

Paleodepth25-50 m@ 417 mcd

MFS/“10”

FS

FS

FS

FS

CumulativeLithology Gamma (cps)

130 140 150 160

%

FS

FS

FS

HS

TTS

T

SystemsTracts

LithologyCore/Recovery

M28, Sequence m5.34

c s vf cSand

Figure 8. Enlargement of the upper part of the transgressive systems tract (TST) and lower highstand systems tract (HST) of the m5.34 sequence at Integrated Ocean Drilling Program Expedition 313 Site M28. Cumulative lithology and lithology columns as in Figure 6; caption and key as in Figure 4. The gamma log (thin purple line) has been smoothed with a 0.5 m fi lter (red line). Arrows point in inferred fi ning direc-tion. Seven fl ooding sequences (FS; parasequence boundaries) and a maximum fl ooding surface (MFS) are inferred by the converging arrows.

Page 11: Testing sequence stratigraphic models by drilling Miocene … · 2013-12-10 · Drilling foresets and testing sequence stratigraphic models Geosphere, October 2013 1237 confusion

Miller et al.

1246 Geosphere, October 2013

to shallowing-upward (generally coarsening upward) facies (Fig. 1) that is recognized using both lithologic and benthic foraminiferal crite-ria. MFS are associated with benthic forami-niferal evidence for deepening upsection to maximum water depths (typically associated with peaks in percent planktonic of total forami-nifera; Loutit et al., 1988) and fi nest grain sizes. Both HST and LST show shallowing-upward successions inferred from coarsening-upward sections and benthic foraminiferal evidence. Transgressive surfaces are generally recognized by a change in stacking pattern from coarsening to fi ning upward (Fig. 1); they are often merged with sequence boundaries on the topsets. TST are transgressive (generally fi ning upward). Parasequence boundaries (fl ooding surfaces) are recognized in LST, TST, and HST by local peaks of percent mud and gamma-ray log stack-ing patterns. We do not identify systems tracts on the bottomsets due to the diffi culty of resolv-ing their complex stratal relationships with the data presented here (Mountain et al., 2010).

RESULTS

Sequence m5.8

Refl ector m5.8 is clearly a seismic sequence boundary, based on onlap, toplap, erosional truncation, and downlap on line 529 (Figs. 3– 5; Supplemental Figs. 22 and 33) and elsewhere in the seismic grids (Monteverde et al., 2008; Monte verde, 2008). A possible FSST underlies the m5.8 seismic sequence boundary at com-mon depth point (cdp) 7900–8100, where there are hints that refl ectors (–2 and 0 in Fig. 5) step down into the basin (offl ap). The overlying m5.7 sequence boundary extensively truncates the topset of the m5.8 sequence landward of Site M27, and the m5.8 sequence is completely eroded ~10 km landward of the site on Line 529. Sequence m5.8 was sampled in the foreset

5010

00

Cum

ulat

ive

Lith

olog

y (%

)

330

320

310

300

290

280

21R

22R

23R

24R

25R

26R

27R

28R

29R

30R

31R

32R

33R

34R

35R

36R

37R

38R

39R

40R

xxxx

xxxx

xxxx

xxxx

xxxxxxxxxxxxxxxx

gg

p

xxx

xx

xx x

xxxx

xxxxxxxxxxxxxxxx

xxxxxxxxxxxxxxxx

xxxxxxxxxxxxxxxx

gxxxxxxxxxxxxxxxx

xxxxxx

xxxx

xxxx

xxxx

xxxx

xxxxxxxxxxxxxxxx

xxxxxx

gg

wd wd

wd w

d

g g

gg

gg

ggLi

thol

ogy

Depth (mcd)

Cor

ere

cove

ryN

GR

(cp

s)

05

1015

TG

R (

cps)

010

020

030

010

Reflector

Depth (mcd)

Age (Ma)

13.7

14.8

15.1

15.7

SO

T

SO

T32

3.23

276.

81

m5.

2/“1

m5/

“14”

10-2

5 m

10-2

5 m

10-2

5 m

10-2

5 m

10-2

5 m

0-10

m

50 m

0-10

m

0-10

m

0-10

m

10-2

5 m

OF

F

Cha

nnel

fill

OF

F

Cha

nnel

fill

Rol

love

rEnvironmentPal

eode

pth

BenthicBiofacies

IntegratedWaterdepth

SystemsTract

25 10 25 50 30 25 1530

HS

T

TS

T?

River-influencedSOT

Seq

uenc

e m

5.2

M27

M28

M29

Tops

etB

ehin

d R

ollo

ver

Fore

set

MF

S

FS

FS

FS

75 m25

m

25-5

0 m

10-5

0 m

OF

F

SF

TS

T

SF

OF

FEnvironmentPal

eode

pth

BenthicBiofacies

IntegratedWaterdepthSystems

35 30 20 3575 10

HS

T

TS

T

Tract

MF

S

76X

77X

78X

79R

80R

81R

82R

83R

84R

Depth (mcd)

220

230

240

Cor

ere

cove

ryLi

thol

ogy

5010

00

Cum

ulat

ive

Lith

olog

y (%

)

m5Reflector

Depth (mcd)

Age (Ma)

m5.

2

225.

45

236.

15

13.7

14.8

15.0

15.6

NG

R (

cps)

05

1015

2025

TG

R (

cps)

020

040

060

0

150

Cum

ulat

ive

Lith

olog

y (%

)

5010

00

Lith

olog

yC

ore

reco

very

151R

152R

153R

154R

155R

156R

157R

158R

159R

160R

161R

162R

600

590

580

126R

127R

128R

129R

130R

131R

132R

133R

134R

135R

136R

137R

138R

139R

140R

141R

142R

143R

144R

145R

146R

147R

148R

149R

150R

570

560

550

540

530

520

510

500

g gg g

gg

gX

gg

g

g gwd g

xxxxxxxxxxxxxxx

xxxxxxxxxxxxxxx

xxxxxxxxxxxxxxx

g

wd

wd

g

gg wd wd

wd wdg wd wd

gwd

wd

Depth (mcd)

OFF

50-6

0 m

(50-

80 m

)50

-60

m (5

0-80

m)

50-8

0 m

75-1

00 m

75-1

00 m

75-1

00 m

75-1

00 m

75-1

00 m

75-1

00 m

75-1

00 m

75-1

00 m

75-1

00 m

75-1

00 m

75-1

00 m

75-1

00 m

75-1

00 m

75-1

00 m

75-1

00 m

75-1

00 m

75-1

00 m

75-1

00 m

75-1

00 m

??25

-50

m

75 m

90-1

00 m

OFF

75-1

00 m

(10-

25?

m)

Environment

Pal

eode

pth

BenthicBiofacies

IntegratedWaterdepth

SystemsTract

75 50 80 90 75 75 75 75

HS

T

TS

T

LST

04

812

NG

R (

cps)

020

040

0

TG

R (

cps)

602.

25m

5.2/

“1”

m5/

“14”

502.

0114

.613

.7

15.8

15.6

Reflector

Depth (mcd)

Age (Ma)

“3”

TS

MF

S

FS

FS

FS

FS

“2”

“3”

“4”

“7”

“8”

“6”

“7”

“8-9

“6”

593

581

cs

vfvc

San

d

cs

vfc

San

d

svf

cS

and

Fig

ure

9. C

ompa

riso

n of

seq

uenc

e m

5.2

at I

nteg

rate

d O

cean

Dri

lling

Pro

gram

Exp

edit

ion

313

Site

s M

27, M

28, a

nd M

29. C

apti

on a

s in

F

igur

e 4.

Pla

cem

ent

of in

tern

al r

efl e

ctor

s 4–

8 in

the

cor

e is

unc

erta

in.

2Supplemental Figure 2. Enlargement of Figure 4. If you are viewing the PDF of this paper or read-ing it offl ine, please visit http://dx.doi.org/10.1130/GES00884.S2 or the full-text article on www.gsapubs.org to view Supplemental Figure 2.

3Supplemental Figure 3. Uninterpreted (top) and interpreted (bottom) seismic profi le Oc270 Line 529 highlighting the m5.8 sequence. Scales are two-way travel-time (TWTT) in seconds and Com-mon Depth Point. Approximate scale in km is given. Dotted line indicates location of Site M27. Arrows indicate refl ector termination. Red are sequence boundaries, blue are transgressive surfaces, green are maximum fl ooding surfaces, and shades of yellow are other refl ectors. Numbers (–3 to 8) are arbitrary designations. If you are viewing the PDF of this paper or reading it offl ine, please visit http://dx.doi.org/10.1130/GES00884.S3 or the full-text article on www.gsapubs.org to view Supplemental Figure 3.

Page 12: Testing sequence stratigraphic models by drilling Miocene … · 2013-12-10 · Drilling foresets and testing sequence stratigraphic models Geosphere, October 2013 1237 confusion

Drilling foresets and testing sequence stratigraphic models

Geosphere, October 2013 1247

at its thickest point (~140 ms, Fig. 5; 133.59 m, Fig. 4) at Site M27; Sites M28 and M29 sam-pled sequence m5.8 in offshore prodelta envi-ronments on the bottomset.

A prominent, high-amplitude refl ector (3 in Fig. 5) onlaps and downlaps the seismic sequence boundary and ties to Site M27 at

~477.52 m composite depth (mcd; Fig. 4). We identify this as the TS at a faint contact zone noted in the core (313-M27–166R-2, 40–56 cm; 477.36–477.52 mcd) based on an uphole change from coarsening upward to fi ning upward at M27 at the level of this refl ector. The LST below this (494.87–477.52 mcd) consists of two

upward-coarsening parasequences (arrows indi-cate fi ning direction, Fig. 4).

Placement of the MFS is unclear in sequence m5.8 at Site M27 (Fig. 4). The TST fi nes upward to at least 460 mcd, with clear coarsening begin-ning above 435 mcd. Lithologic criteria suggest that the MFS occurs in core 158 or 157 where

1(m5.2)

3 4

5 6

7

8

9

(m5.2) 1234567

9

(m5) 14

Onlap

Erosion

Onlap

Toplap Downlap

2

10

1211

13

14(m5)

10111213

1(m5.2)23456789

14(m5)

10111213

M29M28

8

SB

13.7 Ma

TS

SB

TSTMFS

HST

LST

FSST?

Age(Ma)

SystemsTract

14.8 Ma

15.1 Ma15.7 Ma

13.7 Ma14.6 Ma

15.6 Ma15.8 Ma

Onlap Downlap

Sequence m5.2

CDP 5000 4000

0–1

0–1

0-1

Sand and Sandy mud Mud and muddy sand Sand (inferred)

m5.3

Reflector

0 5 km

0.400

0.500

0.600

0.700

TWTT

(sec

)

Figure 10. Interpreted seismic profi le and Wheeler diagram of sequence m5.2 across the foreset at Integrated Ocean Drilling Program Expedition 313 Site M29, extending to the topset at Site M28. Caption as in Figure 5.

Page 13: Testing sequence stratigraphic models by drilling Miocene … · 2013-12-10 · Drilling foresets and testing sequence stratigraphic models Geosphere, October 2013 1237 confusion

Miller et al.

1248 Geosphere, October 2013

mica, laminations, and percent sand reach a minimum; there is no clear observable surface other than a burrowed interval overlying a con-cretion (158–1, 30 cm; 451.36 mcd) that may mark the MFS. Benthic foraminifera indicate deepening upward to the deepest paleodepths at 457.78 mcd where planktonic foraminiferal per-centages peak at 41%. A major downlap surface (5 in Fig. 5, placed at ~442 mcd in Fig. 4) is traced from Sites M28 and M29 (where exten-sive downlap is noted), and carried over the roll-over. It appears to tie to 442 mcd at Site M27. This major downlap surface is the best seismic candidate for an MFS. However, tracing this

surface into the site is unclear and it is possible that the downlap surface correlates deeper (e.g., refl ector 4 in Fig. 5). The slight differences in placement based on seismic, lithologic, and benthic foraminiferal criteria illustrate that pick-ing a defi nitive MFS is not always unequivocal. Our interpretation at Site M27 concludes that the MFS occurs within a zone of maximum fl ooding from 460 to 435 mcd (see Loutit et al., 1988). Above this zone, the HST progressively coarsens upward to fi ne sand at ~415 mcd and above that to a blocky, aggradational medium-coarse sand from 400 mcd to the overlying sequence boundary at 361.28 mcd. Seismic

profi les show a clear progradation from the seis-mic MFS (5 in Fig. 5) to refl ector 7 and general aggradation above this (Fig. 5).

Both Sites M28 and M29 sampled sequence m5.8 in a bottomset location where the dominant facies is tan clayey silt to silty clay deposited in dysoxic prodelta environments (Mountain et al., 2010; Fig. 4). Above the m5.8 sequence boundary at Site M28 (662.98 mcd), there is a thin basal lag of glauconite sand and overly-ing glauco nitic quartz sand, with rapid fi ning upwards to ~660 mcd. The major downlap surface refl ector 5 correlates at 654 mcd to the contact between a silty clay below and uniform

100 150 200 250 300

10 20 30 40 50

m5.2/“1”

m5/“14”

“6”

“8”

“4”

“3”

“2”

Percent plankton0 20 40 60 80 100

%

Dep

th (m

cd)

500

520

540

560

580

600

126R

127R

128R

129R

130R

131R

132R

133R

134R135R136R

137R

138R

139R

140R

141R

142R

143R

144R145R146R

147R

148R

149R

150R

151R

152R

153R

154R

155R

156R

157R

158R

159R

160R

161R

162R

g

gg

g g g

gX

g g

g

g

gwd

g

xxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxx

g

wd

wd

g

gg

wd

wd

wd

wdg

wd

wd

gwd

wd

CumulativeLithology

Gamma (cps) SystemsTracts

LithologyCore/Recovery

MFS

FS

FS

FS

FS

TS

HS

TTS

TLS

T

M29, Sequence m5.2

“7”

s vf cSand

Figure 11. Enlargement of m5.2 sequence in the foreset at Integrated Ocean Drilling Program Expedition 313 Site M29. Lithology and cumulative lithology columns as in Figure 9 (see Fig. 4 caption). Percent plankton of total foraminifera indicated in magenta. The gamma log (thin black line) has been smoothed with a 0.6 m fi lter (red line). Arrows point in inferred fi ning direction.

Page 14: Testing sequence stratigraphic models by drilling Miocene … · 2013-12-10 · Drilling foresets and testing sequence stratigraphic models Geosphere, October 2013 1237 confusion

Drilling foresets and testing sequence stratigraphic models

Geosphere, October 2013 1249

prodelta clayey silt above, suggesting that this is the deepwater equivalent to the MFS. Benthic foraminifera are absent from the m5.8 sequence at Site M28. At Site M29, glauconitic siltstones deposited in offshore environments overlie the sequence boundary (753.80 mcd). The seismic downlap surface (refl ector 5) at Site M29 cor-relates with an upward change to uniform pro-delta clayey silts. Benthic foraminifera indicate paleodepths of 50–80 m immediately above the sequence boundary; paleodepths increase upsec-tion to 75–100 m, and possibly decrease to 50–100 m at the top of the sequence. It is not pos-sible to defi nitely assign these bottomset deposits below the deepwater equivalent to the MFS at Sites M28 and M29 to the LST or TST based on seismic, lithologic, or benthic foraminiferal crite-ria, although at least some equivalence to the TST at Site M27 is implied (see correlations in Fig. 4).

Sequence m5.8 appears to be a million-year–scale sequence based on seismic, lithologic, benthic foraminiferal, and age criteria. The Wheeler diagram (Fig. 5, top) also suggests that it is one sequence. The m5.8 sequence is dated as 20.1–19.2 Ma at Site M27 in the foreset and as 20.0–19.5 Ma at Site M28 and 20.2–20.0 Ma at Site M29 in the bottomsets, suggesting that the bottomsets do not record the younger part of the sequence (Fig. 4; Browning et al., 2013). The basal m5.8 sequence boundary correlates with the Miocene oxygen isotope event Mi1aa δ18O increase based on biostratigraphy and sta-ble isotope stratigraphy (Browning et al., 2013), a relatively minor glacioeustatic lowering (i.e., 0.8‰ increase corresponding to ~40 m lower-ing). It also correlates with the Burdigalian-1 sequence boundary of ExxonMobil (Snedden and Liu, 2010).

Sequence m5.4 Composite Sequence

Site M28 was designed to sample the thickest part of sequence m5.4 on the foreset, close to the rollover of the overlying m5.3 sequence bound-ary (Figs. 3, 6, and 7; Supplemental Figs. 44 and 55; Mountain et al., 2010). On line 529 (Fig. 7), the sequence is bracketed by two high-ampli-tude, prominent refl ectors (m5.4 and m5.3; Figs. 3 and 7) associated with onlaps, downlaps (e.g., refl ector 5 in Fig. 7), toplaps, and erosional

truncations. These are clear seismic sequence boundaries and they have been traced through the seismic grid (Monteverde et al., 2008; Monte verde, 2008).

Refl ections 0 to 3 (Fig. 7) underlying the m5.4 seismic sequence boundary are part of the underlying m5.45 sequence (Fig. 7) and may represent an FSST because they appear to step down, although this may merely be a result of truncation of the HST by the m5.4 sequence boundary. Tracing sequence boundary m5.4 and distinguishing it from the possible FSST is clear if criteria of onlap, downlap, erosional trunca-tion, and toplap are followed.

At the million-year scale, sequence m5.4 is interpreted seismically to consist of (1) a thick LST (123 m) evidenced by weak aggradation to refl ector m5.34 (7) and strong prograda-tion above m5.34 to the major downlap surface marked by refl ector m5.32 (14) (Figs. 3 and 7), and (2) a 30-m-thick progradational to aggrada-tional HST above the m5.32 downlap surface to the overlying m5.3 sequence boundary. There apparently is no seismic evidence for an inter-vening TST (Fig. 7). However, the million-year-scale sequence m5.4 (spanning ca. 17.7–16.7 Ma at Site M28; Fig. 6) is a composite sequence (sensu Mitchum and Van Wagoner, 1991; Neal and Abreu, 2009; Flint et al., 2011) that can be parsed into three sequences, m5.4–1, m5.34, and 5.33 (we use the term 5.4–1 to differentiate the higher frequency sequence, but both the million year and higher frequency sequences share the same basal sequence boundary, refl ector m5.4). Coring and logging reveal that this sequence has a very complex internal structure, and integration of seismic, lithologic, foraminiferal, and log cri-teria justify recognizing three distinct sequences within the m5.4 composite sequence.

Lithologic and benthic foraminiferal pat-terns are key criteria to resolving this composite sequence (Fig. 6). Two coarsening-upward para-sequences separated by a thin fi ning-upward succession occur at Site M28 between the m5.4 sequence boundary (512.33 mcd) and refl ec-tor 5 (Figs. 6 and 7). This 11-m-thick interval is interpreted as the LST. Refl ector 5 (Fig. 7) correlates to a level where there is a change from coarsening to fi ning upward in the cores at 501 mcd, and is thus interpreted as a TS

(Fig. 6). The LST is overlain by an abruptly fi ning-upward succession from 501 to 494 mcd that is interpreted as the TST (Fig. 6). Benthic foraminiferal bio facies, percent plankton, and grain size changes all indicate deepening in the TST above 501 mcd to an MFS associated with refl ector 6 at 494 mcd (Fig. 6). The section then coarsens upsection in the HST to a major refl ec-tor (7, m5.34) at 479 mcd (Figs. 6 and 7).

We interpret m5.34 as a seismic and core sequence boundary. It shows onlap by refl ec-tors 8 and 10, downlap by refl ectors 8 and 9, and erosionally truncates the m5.4 sequence boundary (Fig. 4). We traced m5.34 to adjacent profi les in the seismic grid and found evidence that it is a seismic sequence boundary using cri-teria of onlap, downlap, erosional truncation, and toplap.

Lithologic, foraminiferal, and log data can be used to recognize systems tracts within the m5.34 sequence (Fig. 6). At Site M28, there is a coarsening-upward succession immedi-ately above m5.34 (479 mcd) to ~475 mcd that we interpret as an LST (Fig. 6). The latter is approxi mately the level of refl ector 8 (467 mcd) that downlaps and onlaps m5.34 (Fig. 7). Thus, we interpret refl ector 8 as a TS, and suggest its correlation at 475 mcd, 8 m below its predicted depth. Subsequent fi ning upward occurs from ~475 to ~468 mcd (Fig. 6) in the lower part of the TST. It is diffi cult to pick the MFS for the m5.34 sequence because the section lacks foraminifera below 430 mcd (presumably due to dissolution), the cumulative lithology is com-plicated by the interlaminations of sand and silt that obscure trends, and the dynamic range of the gamma-log values (Fig. 6) is dampened by larger variations above and below.

Examining parasequences within the m5.34 sequence at Site M28 allows us to identify the MFS. Expanding the gamma log (Fig. 8) shows values increasing from 470 to 449 mcd (punctuated by decreases at ~466, ~460, and ~454 mcd), and then generally decreasing to 417 mcd, where there is an abrupt shift to low gamma-log values (Fig. 6). We interpret this as four progressively deeper parasequences, with the MFS identifi ed by gamma logs at 449 mcd in a coring gap (Fig. 8); lithologic descriptions similarly note the change from fi ning to coars-ening upward at ~445 mcd (Mountain et al., 2010). A downlap surface (refl ector 10, Fig. 7) correlates to Site M28 at ~449 mcd, suggest-ing that this is the MFS. The sequence coarsens upsection in the HST (445–405 mcd) and ben-thic foraminifera show evidence for shallowing. Decreasing gamma-log values upsection are consistent with coarsening upward, with 5 pro-gressively shallower parasequences indicated by FS at 442, 435, 432, and 427 mcd (Fig. 8).

4Supplemental Figure 4. Enlargement of Figure 6. If you are viewing the PDF of this paper or reading it offl ine, please visit http://dx.doi.org/10.1130/GES00884.S4 or the full-text article on www.gsapubs.org to view Supplemental Figure 4.

5Supplemental Figure 5. Uninterpreted (top) and interpreted (bottom) seismic profi le Oc270 Line 529 high-lighting the m5.4 composite sequence. Scales are two-way travel-time (TWTT) in seconds and Common Depth Point. Approximate scale in km is given. Vertical red line indicates location of Site M28. Arrows indicate refl ector termination. Red are sequence boundaries, blue are transgressive surfaces, green are maximum fl ood-ing surfaces, and shades of yellow are other refl ectors. Numbers (0 to 21) are arbitrary designations. If you are viewing the PDF of this paper or reading it offl ine, please visit http://dx.doi.org/10.1130/GES00884.S5 or the full-text article on www.gsapubs.org to view Supplemental Figure 5.

Page 15: Testing sequence stratigraphic models by drilling Miocene … · 2013-12-10 · Drilling foresets and testing sequence stratigraphic models Geosphere, October 2013 1237 confusion

Miller et al.

1250 Geosphere, October 2013

The parasequences in the TST have thicker fi ning-upward successions overlain by thinner coarsening-upward successions; in the HST, the pattern is reversed, with thinner fi ning-upward and thicker coarsening-upward successions.

We tentatively interpret refl ector m5.33 as a sequence boundary based on onlap and down-lap, and due to the major downlap onto refl ector m5.32 (refl ector 14), we interpret it as an MFS. The absence of intersecting profi les with clear seismic defi nition means that loop correlations cannot confi rm that m5.33 is a seismic sequence boundary. At Site M28, candidate sequence boundary m5.33 correlates to ~405 mcd in an interval of poor recovery. A change at 393 mcd from a coarsening- to a fi ning-upward succes-sion marks the change from the LST to a TST and placement of the TS at this level. Refl ector m5.32 (refl ector 14) correlates to 391 mcd at a large gamma kick associated with a change from fi ning upward to coarsening upward. Benthic foraminiferal evidence and percent planktonic foraminiferal evidence indicate a maximum paleowater depth within this sequence at the level of this MFS. Coarsening associated with progradation continues from 391 mcd upward to 380 mcd, ending with blocky, aggradational sands at the top. The HST above m5.32 at Site M27 is seismically composed of a series of inclined and downstepping refl ectors possibly refl ecting an FSST or erosional truncation of the HST clinoforms (Fig. 7).

The age-distance Wheeler diagram clearly illustrates the nature of the composite sequence (Fig. 7). The m5.4–1 sequence steps seaward of the previous m5.45 sequence and then steps landward, but is truncated by the overlying m5.34 sequence, with its HST poorly devel-oped. The m5.34 sequence steps farther sea-ward than the underlying sequence, and then fully landward in the TST, with a better devel-oped HST. The m5.33 sequence steps farther seaward than m5.4–1 and m5.34, with the best developed HST. Overall m5.4–1 and m5.34 are progradational and m5.33 is aggradational to progradational. We note that lower resolution seismic data and/or poor core recovery would most likely have failed to resolve each of these embedded sequences, and the composite m5.4 sequence would have been interpreted as a thick LST (which in reality is the m5.4–1 and m5.34 sequences and LST of m5.33) with a thinner, highly downlapping HST (which is the HST of the m5.33 sequence).

Site M27 sampled the million-year-scale m5.4 sequence at a topset where it is composed of the m5.34 and m5.33 sequences; the m5.4–1 sequence appears to have been eroded at this location (Fig. 7). The m5.34 sequence consists of a thin transgressive lag above the sequence

boundary (295.01 mcd) and a thin TST that fi nes up to an MFS at 288 mcd. The HST coars-ens upsection to the m5.33 sequence boundary (271.23 mcd) and is thus 17 m thick. In the m5.33 sequence, a possible thin TST (271.23–265 mcd) is overlain by an especially mud-rich interval with the deepest paleodepth within this sequence, based on benthic forami nifera, strongly suggesting an MFS at ~265 mcd. A thin (~9 m) HST caps the sequence, ending at the overlying m5.3 sequence boundary (preferred placement at 256.19 mcd, although it could be placed at 249.75 mcd; see Miller et al., 2013). Thus, both sequences 5.34 and m5.33 at Site M27 consist of thin TST and moderately thick HST on the topsets. Based on lithology the m5.33 sequence is fi ner grained at Site M27 than at the more basinward Site M28. Fur-thermore, water depth estimates for m5.33 are deeper at Site M27 than at M28. We interpret this as indicating that the m5.33 sequence at Site M27 represents only the upper TST and lower HST and that this same interval is expressed as a hiatus (0.7 m.y.) at Site M28.

Composite sequence m5.4 was sampled at Site M29 in a bottomset setting and dated as 17.7–17.6 Ma (Figs. 6 and 12). This suggests that the bottomset portion correlates with the m5.4–1 sequence, although the age resolu-tion allows correlation to the m5.34 sequence. Seismic correlations suggest that the m5.34 sequence is present at Site M29. The bottomset consists of fairly uniform silts with transported glauconite sandstone beds.

Age estimates for the m5.4-m5.34-m5.33 composite sequence are consistent with more than one sequence. Sr isotope age estimates show a mean linear fi t of 17.7–16.7 Ma for the m5.4 composite sequence at Site M28. In Browning et al. (2013), the ages of m5.4–1 (17.75–17.67 Ma), m5.34 (17.60–17.40 Ma), and m5.33 (16.70–16.60 Ma) were estimated. Maximum theoretical resolution for this portion of the Sr isotope curve is ±0.3 m.y. (see discus-sion in Browning et al., 2013). Given this, the mean age of m5.33 (16.65 Ma) is statistically different from the older two ages, although the mean ages of m5.34 (17.5 Ma) and m5.4–1 (17.65 Ma) are not statistically different. Thus, it is clear that the age control requires at least two sequences with a signifi cant hiatus sepa-rating them.

The basal sequence boundary of the composite sequence m5.4 (ca. 17.7 Ma) correlates with the Mi1b δ18O increase (17.7 Ma; Browning et al., 2013), a relatively minor glacioeustatic lowering (i.e., ~0.8‰ increase corresponding to ~40 m lowering). It also correlates with the Burdi-galian-4 sequence boundary of ExxonMobil (Snedden and Liu, 2010). The correlation of the

m5.34 and m5.33 sequence boundaries to δ18O variations is uncertain due to the lack of high-resolution data in this interval, although the hia-tus between m5.4 and m5.34 (17.4–16.7 Ma) may correlate with a 400-k.y.-scale increase ca. 16.8 Ma. Deposition of the m5.33 sequence cor-relates with an interval of peak sea level in the early Miocene climatic optimum (Fig. 12).

Sequence m5.2

The basal m5.2 sequence boundary is defi ned by onlap, downlap, erosional trunca-tion, and toplap on line 529 (Figs. 3, 9, and 10; Supplemental Figs. 66 and 77) and elsewhere in the available seismic grid. A possible FSST occurs below the sequence boundary in the m5.3 sequence (refl ectors –1, 0; cdp 4900–4950, Fig. 10), although this could be due to trunca-tion of the HST of the underlying sequence by m5.2. The m5.2 basal sequence boundary cor-relates to 602.25 mcd at Site M29, where it was sampled in the lower foreset (Fig. 9). Refl ector 2 in Figure 10 onlaps and downlaps the m5.2 sequence boundary and correlates to 593 mcd at Site M29; this is immediately above the top of a coarsening-upward succession at ~593 mcd, suggesting that the TS is at 593 mcd and that the LST is ~9 m thick. The overlying TST (~593–581 mcd) fi nes upsection and is capped by a prominent downlap surface (3) at ~581 mcd interpreted as the MFS. High planktonic foraminiferal abundances at 576.76 (Fig. 11) support placement of the MFS near refl ector 3. A thick (79 m) HST above this contains several FS within it (Figs. 9, 10, and 11), consistent with the presence of at least 4 downlap surfaces noted on the seismic profi le (Fig. 10), refl ec-tors 3 (the MFS), 4, 5, and 8. Downlap is not obvious on seismic refl ectors 6 and 7. However, we note that refl ectors 4, 6, 7, and 8 correlate with fl ooding surfaces noted in the gamma logs and lithology as mud peaks (Figs. 9 and 11);

6Supplemental Figure 6. Enlargement of Figure 9. If you are viewing the PDF of this paper or read-ing it offl ine, please visit http://dx.doi.org/10.1130/GES00884.S6 or the full-text article on www.gsapubs.org to view Supplemental Figure 6.

7Supplemental Figure 7. Uninterpreted (top) and interpreted (bottom) seismic profi le Oc270 Line 529 highlighting the m5.2 sequence. Scales are two-way travel-time (TWTT) in seconds and Common Depth Point. Approximate scale in km is given. Vertical red line indicates location of Site M29. Arrows indicate refl ector terminations. Red are sequence bound aries, blue are transgressive surfaces, green are maxi-mum fl ooding surfaces, and shades of yellow are other refl ectors. Numbers (–1 to 14) are arbitrary designations. If you are viewing the PDF of this paper or reading it offl ine, please visit http://dx.doi.org/10.1130/GES00884.S7 or the full-text article on www.gsapubs.org to view Supplemental Figure 7.

Page 16: Testing sequence stratigraphic models by drilling Miocene … · 2013-12-10 · Drilling foresets and testing sequence stratigraphic models Geosphere, October 2013 1237 confusion

Drilling foresets and testing sequence stratigraphic models

Geosphere, October 2013 1251

the slight offset in depths (2–4 m) appears to be consistent and due to a minor problem with the velocity-depth function. Onlap onto refl ec-tors 3 and 8 suggests that they may be sequence boundaries and that m5.2 is also a composite sequence. We lack the data to make this inter-pretation, although erosional surfaces noted in the cores at 577.89 and 573.66 mcd may be a higher frequency sequence boundary and TS. It is possible that the downstepping associated with refl ectors 9–13 represents an FSST (Fig. 10), although erosional truncation of this sec-tion could also explain apparent downstepping.

At Site M28, sequence m5.2 was sampled immediately landward of the rollover where it consists of a thin TST and a thick HST with three FS indicated by mud and gamma-log peaks (Fig. 9) and seismic downlap surfaces 6, 7, and 8-9 (Fig. 10). At Site M27, sequence m5.2 consists of a thin (~6 m) TST and thin HST sampled on a topset (Fig. 9).

The age-distance Wheeler diagram (Fig. 10) shows that the m5.2 sequence is predominantly aggradational, although immediately above refl ector 8 it becomes strongly progradational to refl ector 10, where it apparently steps seaward and downward as a possible FSST (refl ectors 10–13). Foreset beds of m5.2 at Site M29 (where the sequence is thickest) are ca. 15.6–14.6 Ma (Fig. 12). Rollover (Site M28) and topset (Site M27) strata are 15.1–14.8 Ma, suggesting non-deposition of the LST and lower TST and the upper HST (Fig. 9). The basal m5.2 sequence boundary (15.6 Ma) appears to be younger than the major Mi2 δ18O increase (16.3 Ma) and older than the major Mi2a (14.6 Ma) (both >1‰, >50 m eustatic fall). It may be associ-ated with a smaller (0.8‰, ~40 m eustatic fall) 400-k.y.-scale δ18O increase (Fig. 12), although age control in this interval is less certain and it is possible that it correlates with Mi2a within the age constraints. We suggest it correlates with the Bur5-Lan1 sequence boundary of Exxon-Mobil (16 Ma; Snedden and Liu, 2010).

DISCUSSION

Systems Tracts and Sequence Stratigraphic Models

Our systems tracts interpretations allow us to test sequence stratigraphic models, particu-larly in the foresets where we recovered low-stand deposits. Drilling through the foresets yields generally thin LST (<18, 11, 4, 12, and 9 m thick for sequences m5.8, m5.4–1, m5.34, m5.33, and m5.2, respectively; Figs. 4, 6, and 9). On the foresets, we also identifi ed thin TST (26, 7, 26, 2, 12 m thicknesses for sequences m5.8, m5.4–1, m5.34, m5.33, and m5.2, respec-

tively). However, thick HST occur on the fore-sets (90, 15, 44, 30, and 79 m thicknesses for sequences m5.8, m5.4–1, m5.34, m5.33, and m5.2, respectively; Figs. 4, 6, and 9). LST on the foresets consist of one (Fig. 9) to two (Figs. 4 and 6) coarsening-upward parasequences. TS are recognized in foresets by shifts from coars-ening-upward successions to fi ning-upward successions. TST on the foresets record para-sequences as overall thick fi ning-upward (deep-ening) successions punctuated by thin coarsen-ing-upward (shallowing) parasequences (e.g., Figs. 8 and 11). HST on the foresets refl ect the inverse, because thin fi ne-grained units overlie thicker coarsening-upward parasequences (Figs. 8 and 11).

Topsets consist of shallow-water deposits (shoreface to middle neritic) above merged surfaces that represent both TS and sequence boundaries. TST on topsets consist of fi ning- and deepening-upward successions overlain by coarsening- and shallowing-upward HST.

Bottomsets consist of downslope-transported sands and hemipelagic muds deposited in 75–100 m water depths (Mountain et al., 2010). Facies successions within bottomsets are not discussed here.

FSST are possibly recognized below seismic sequence boundaries below the rollover. Exam-ples are shown on line 529 in sequence m5.45 below sequence m5.4 (Fig. 7), in m5.3 below m5.2 (Fig. 10), and possibly in m6 below m5.8 (Fig. 5). These FSST have not been confi rmed on adjacent profi les. Where sampled, these pos-sible FSST appear to consist of blocky sands (Figs. 7 and 10).

Our interpretation of thin LST contrasts with published seismic stratigraphic predictions of thick LST and thin to absent TST. Greenlee et al. (1992) examined widely spaced profi les tied to logs of exploration wells and proposed that Miocene sequences on the New Jersey shelf stratigraphically above our sequence m5 (their “Green” sequence) were dominated by LST. In Monteverde et al. (2008) and Monteverde (2008), thick LST and thick HST for sequences m5.8, m5.4, and m5.2 were also interpreted (Fig. 13). Here we compare these former inter-pretations with our conclusions that have the benefi t of higher resolution and more densely spaced seismic data, along with core and log integration. Interpretations based on seismic profi les alone (Fig. 13, bottom) tend to over-estimate the extent and thickness of LST while underestimating TST (Fig. 13). Possible reasons why LST are overestimated include the follow-ing. (1) TS are diffi cult to distinguish seismi-cally; this explains the different interpretations of sequence m5.8 (Fig. 13). (2) Composite sequences can contain stacked higher frequency

sequences that are diffi cult to distinguish from LST; this explains the different interpretations of sequence m5.4 (see following for further dis-cussion). (3) Sequences contain multiple down-lap surfaces, the stratigraphically lowest being the MFS; this explains the different interpreta-tions of sequence m5.2.

We fi nd no evidence for sequence boundaries expressed as correlative conformities in the shal-low (<120 m paleodepth) sequences sampled by Expedition 313. We show on the age-distance Wheeler diagrams that in the foresets, where sequences are supposed to be most complete, there is evidence of erosion (Figs. 5, 7, and 10) and hiatuses. For example, we note hiatuses of 0.6, 0.2, 0.7, and 0.2 m.y. associated with the m5.8, m5.4-1, m5.33, and m5.2 sequence boundaries in the foresets, respectively. Longer hiatuses occur on the bottomset, presumably due to erosion and sediment bypass associated with downslope processes (Mountain et al., 2010). Only the higher frequency sequence bound-ary m5.34 has no discernible hiatus and may refl ect continuous deposition (Fig. 7). There are several other sequence boundaries with no dis-cernible time gaps with the resolution available (0.25–0.5 m.y.) (Browning et al., 2013); how-ever, there is still core evidence of erosion in the cores associated with sequence boundaries, even in bottomsets.

If the correlative conformity exists, it is on the continental slope, but even there, hiatuses are associated with sequence boundaries and downslope transport (Miller et al., 1996). ODP Site 904 (Mountain et al., 1996) drilled Mio-cene sequences on the slope (1123 m water depth) where a long hiatus (15.6–13.6 Ma) encompassing the m5.2 sequence described here was reported (Miller et al., 1996), plus short hiatuses (16.9–16.3 Ma, ca. 22–21 Ma) and inferred continuous sedimentation from 21 to 16.9 Ma encompassing sequences m5.8 to m5.6 described here. However, sedimentation rates in the interval of inferred continuity on the slope are low (~10 m/m.y.) and continuous sedi-mentation is unproven. Reevaluation of correla-tions to Site 904 and the chronology there will be the subject of future work. In Mountain et al. (2007), Pleistocene refl ectors were traced to the New Jersey continental slope ODP Site 1073 (650.9 m water depth), where continuity is dem-onstrated by correlation to δ18O records on the Milankovitch scale; two sequence boundaries in particular, p2 and p3, correlate with marine isotope chrons 8–9 (300 ka) and 11–12 (424 ka), respectively, and exhibit no obvious hiatuses. In contrast, Aubry (1993) found no evidence of continuity for Miocene slope sequences in the Desoto Canyon area (west Florida). Studies of a corehole on the continental slope (300 m) in

Page 17: Testing sequence stratigraphic models by drilling Miocene … · 2013-12-10 · Drilling foresets and testing sequence stratigraphic models Geosphere, October 2013 1237 confusion

Miller et al.

1252 Geosphere, October 2013

2728

29E

xped

ition

313

Miocene

DN7 6 DN5b DN4 DN3 2c 2b DN2a DN1

Serravallian Langhian Burdigalian Aquitanian

middle early

5A AA AB AC AD 5B 5C 5D 5E 6 6A 6AA 6B 6C 7

M9b 9a 8 M7 M6 M5b 5a 4b M4a M3 M2 M1b M1a O7

NN6 NN5 NN4 NN3 NN2 NN1

m5.

2

m5.

3

m5

m5.

8

m5.

45

m5.

3

m5.

2

m5.

7m

5.47

/m5.

6m

5.47

m5.

45m

5.4

m5.

6

m5.

3

m5.

2

m4.

5

m5.

47/m

5.7

4.2-

4.5

mer

ged

m5

m5.

45

m5.

4

Glo

balδ

18O

Spl

ice

‰ V

PD

B (

adju

sted

to e

quili

briu

m)

11.

52

2.5

33.

5

15 20 2512 13 14 16 17 18 19 21 22 23 24

m6

m5.

7

m4.

4/3/

2

m5

m4.

5

m5.

8

7 6b 6a 5 4 3 2 ECDZ1aDiatoms

DinocystsCalcareousnannoplanktonForaminifers

Polarity

Epoch

Age

Age (Ma)

Chron

0.5

11.

52

2.5

3

Mi1

Mi2

a

Mi3

Mi4

Mi1

a

Mi1

b

Mi2

Mi1

aa

Mi1

ab

Mi5

m4.

1

m3*

m4*

m2*

m1*

100

41 100

41?

41?

100

100

15 20 2512 13 14 16 17 18 19 21 22 23 24

m6

unna

med

Dominantperiodicity

*dat

ed a

t slo

pe S

ite 9

04

14.6

16.3

17.7

18.7

12.0

13.0

13.7

15.1

15.8

18.4

19.2

19.6

20.3

5

20.8

21.2

22.6

23.2

20.0

Site 1090Site 926 and 929Site 574Site 574 588

1.5

2.5

33.

54

2

?

Oligocene

1b

5a

m5.

4m

5.34

m5.

33

22.0

22.3

18.0

17.4

16.8

m5.

8

41

O6

m6

m5.

33

Sed

imen

tatio

nra

tes

atS

ite M

29(m

/Myr

)

96100

236

160

0

Fig

ure

12. A

ges

of s

eque

nces

m6

to m

5 an

d co

rrel

atio

n to

dee

p-se

a δ18

O r

ecor

ds (

mod

ifi ed

aft

er B

row

ning

et

al.,

2013

). P

ink

hori

zont

al

indi

cate

s in

fl ect

ions

in

δ18O

inc

reas

es.

Col

ored

blo

cks

show

tim

e pe

riod

s re

pres

ente

d by

Int

egra

ted

Oce

an D

rilli

ng P

rogr

am E

xped

itio

n 31

3 co

reho

les

com

pare

d to

the

glob

al o

xyge

n is

otop

e re

cord

(VP

DB

—V

ienn

a P

eede

e be

lem

nite

). M

i pre

fi x is

Mio

cene

oxy

gen

isot

ope

even

t.

Tim

e sc

ale

of G

rads

tein

et

al.

(200

4).

Cor

ehol

es a

rran

ged

from

upd

ip o

n le

ft t

o do

wnd

ip o

n ri

ght.

Dif

fere

nt c

olor

ed b

lock

s ar

e us

ed t

o sh

ow w

here

the

sit

e is

loca

ted

wit

h re

spec

t to

the

clin

othe

m a

nd t

he t

ime

repr

esen

ted

by a

giv

en s

eque

nce:

ligh

t bl

ue (

tops

et),

med

ium

blu

e (f

ores

et),

and

dar

k (b

otto

mse

t). C

ross

-hac

hure

s in

dica

te le

ss c

erta

in a

ge t

uned

to

bett

er d

ated

rec

ords

at

othe

r si

tes.

Ben

thic

for

amin

ifer

al

oxyg

en is

otop

e re

cord

s fr

om S

ite

574

(blu

e), S

ite

588

(blu

e), S

ite

1090

(gre

en),

and

Sit

e 92

6 (b

lack

). T

he d

omin

ant p

erio

ds fo

und

in d

eep-

sea

δ18O

rec

ords

are

sum

mar

ized

aft

er P

älik

e et

al.

(200

6) a

nd H

olbo

urn

et a

l. (2

007)

, and

are

indi

cate

d by

ora

nge

(ecc

entr

icit

y do

min

ated

, 100

an

d 40

5 k.

y.)

vers

us v

iole

t (t

ilt d

omin

ated

, 41

k.y.

) Se

e B

row

ning

et

al. (

2013

) fo

r ci

tati

ons.

Page 18: Testing sequence stratigraphic models by drilling Miocene … · 2013-12-10 · Drilling foresets and testing sequence stratigraphic models Geosphere, October 2013 1237 confusion

Drilling foresets and testing sequence stratigraphic models

Geosphere, October 2013 1253

7000

6000

5000

Shot

poi

nt

0.40

0

0.50

0

0.60

0

TWTT (sec)

Sequ

ence

m5.

4S

eism

ic-c

ore-

log

inte

rpre

tatio

n

M27

M28

Sho

t poi

nt50

0040

00

0.40

0

0.50

0

0.60

0

0.70

0

TWTT (sec)

M28

M29

Seq

uenc

e m

5.2

Sei

smic

-cor

e-lo

g in

terp

reta

tion

NW

SE

0.50

0

1.00

0

3500

3000

4000

4500

5000

5500

6000

6500

7000

7500

8000

8500

9000

Seq

uenc

e m

5.2

Seq

uenc

e m

5.4

Seq

uenc

e m

5.8

2500

0.40

0

0.50

0

0.60

0

0.70

0

TWTT (sec)

8000

7000

Seq

uenc

e m

5.8

Sei

smic

-cor

e-lo

g in

terp

reta

tion

M27

Shot

poi

nt

Sei

smic

inte

rpre

tatio

nLi

neO

C27

0 di

p lin

e 52

9

M27

M28

M29

Shot

poi

nt

TWTT (sec)

Low

stan

d sy

stem

s tra

ct

Tran

sgre

ssiv

e sy

stem

s tra

ct

Hig

hsta

nd s

yste

ms

tract

Falli

ng s

tage

sys

tem

s tra

ct?

m5.

47m

5.7

m5

m6

m5.

2m

5.3

m5.

4m

5.45

m5.

8

10 k

m0

Sca

le

?

?

?

Fig

ure

13.

Com

pari

son

of s

yste

ms

trac

ts p

redi

cted

fro

m s

eism

ic p

rofi l

es (

belo

w;

Mon

teve

rde

et a

l., 2

008;

Mon

teve

rde,

200

8) w

ith

the

seis

mic

-cor

e-lo

g in

terp

reta

tion

s de

velo

ped

here

(to

p 3

pane

ls).

Sec

tion

s ar

e in

tw

o-w

ay t

rave

l-ti

me

(TW

TT

) in

sec

ond

(s)

vers

us c

dp (

com

-m

on d

epth

poi

nt);

app

roxi

mat

e sc

ale

(in

km) i

s sh

own.

Loc

atio

ns o

f sit

es a

re s

how

n by

ver

tica

l red

line

s an

d ap

prox

imat

e de

pth

is in

dica

ted

usin

g a

velo

city

-dep

th f

unct

ion.

Bas

al s

eque

nce

boun

dari

es d

iscu

ssed

her

e ar

e in

red

.

Page 19: Testing sequence stratigraphic models by drilling Miocene … · 2013-12-10 · Drilling foresets and testing sequence stratigraphic models Geosphere, October 2013 1237 confusion

Miller et al.

1254 Geosphere, October 2013

the Gulf of Lion (Western Mediterranean Sea) show very expanded glacial sections and con-densed interglacial sections (Sierro et al., 2009), contradicting previous seismic interpretations of refl ectors as correlative conformities corre-sponding to the low sea levels caused by gla-cial buildup. We conclude that the existence of a correlative conformity is unproven and should not be considered a cornerstone of sequence stratigraphy.

Higher Frequency Sequences and Sequence Hierarchy

We agree with many studies that recognize that sequences on the million-year scale can be the composite of smaller scale sequences (e.g., Mitchum and Van Wagoner, 1991; Neal and Abreu, 2009; Flint et al., 2011). Here we show that sequence m5.4 is a composite sequence comprising three higher frequency sequences. The composite m5.4 sequence shows a change from a thick aggradation-progradation succes-sion to extensively progradational succession across a major downlap surface (m5.32); on the million-year scale, this would be interpreted as dominantly LST, no TST, and a thin HST. How-ever, we show that the LST are actually very thin within the three sequences that comprise composite sequence m5.4. This is illustrated by Figure 13, which shows the million-year-scale interpretation based on seismic interpretations (bottom panel) versus the integrated interpreta-tion that requires three sequences (top panel). We suspect that there is additional detail still to be detected within sequence m5.2 as well, and it may also be composite, but available data are insuffi cient to evaluate this. This underscores the long-recognized fact that the ability to resolve sequences depends on seismic resolu-tion. Sequences fi ner than the million year scale can be usually be resolved only in regions with high accommodation and sediment supply (e.g., Abdulah and Anderson, 1994), with very high resolution seismic data, or from detailed outcrop mapping over large areas (e.g., DiCelma et al., 2011; Flint et al., 2011).

There have been two approaches to classify-ing sequence hierarchy. The Exxon approach has been to recognize hierarchical orders of sequences, with fi rst order (108 yr scale) due to tectonism, second order (107 yr) and third order (106 yr scale) due to various possible processes, and higher order scales due to Milankovitch forcing on the 405 k.y., 100 k.y., 41 k.y., 23 k.y., and 19 k.y. scales (Vail et al., 1977; Mitchum and Van Wagoner, 1991). Schlager (2004) sug-gested that sequences and systems tracts are scale-invariant fractal features and that they do not follow hierarchical orders. Boulila et al.

(2011) noted that icehouse (Oligocene to Holo-cene) million-year-scale δ18O variations were paced by the 1.2 m.y. tilt cycle; they suggested that sequences appear to follow the 1.2 m.y. cycle due to glacioeustatic forcing. In contrast, greenhouse (Cretaceous–Eocene) sequences seem to be paced by the 2.4 m.y. eccentricity cycle, although this has not been demonstrated unequivocally.

Oxygen isotope studies show that although million-year-scale ice volume variability was dominated by the 1.2 m.y. tilt cycle, there were numerous changes in the dominant higher frequency pacemaker in the early to middle Miocene, from eccentricity (100 and 405 k.y.) dominated to tilt (41 k.y.) dominated benthic foraminiferal δ18O variations (Pälike et al., 2006; Holbourn et al., 2007). Sequences m5.8 and m5.2 were deposited in a 100 k.y. cycle–domi-nated world, indicated by wavelet analysis of δ18O data (Pälike et al., 2006) (Fig. 12). Unfor-tunately, δ18O resolution is insuffi cient at present to document the dominant pacing of the interval from 18.5 to 16.6 Ma, the time encom passing composite sequence m5.4 (Fig. 12). Higher fre-quency sequences within the m5.4 composite sequence suggest response to the 100 and/or 400 k.y. eccentricity cycles and perhaps even the 23 and 19 k.y. precessional cycles (Fig. 12).

Our chronology is consistent with oxygen isotope studies indicating that early Miocene sequences were paced by 1.2 m.y. tilt and 100 k.y. and 405 k.y. eccentricity cycles (Fig. 12). Sequence m5.8, composite sequence m5.4, and sequence m5.2 have been dated (20.1–19.2, 17.7–16.6, and 15.6–14.6 Ma; Browning et al., 2013) with durations of 0.9, 1.1, and 1 m.y., respectively, close to the 1.2 m.y. predicted by Milankovitch glacioeustatic forcing (Fig. 12). The 3 sequences and hiatuses within the m5.4 composite sequence constrain the duration of the sequences to 400 k.y. or shorter time scales. Our age model suggests durations of ~80, ~200, and ~100 k.y. for the 3 higher frequency sequences m5.4–1, m5.34, and m5.33. However, age con-trol is no better than ±250 k.y., and thus we cannot demonstrate that these sequences were forced by the 100 k.y. or the longer 405 k.y. eccentricity cycle. Nevertheless, log data pro-vide intriguing hints of much higher resolution forcing that may be a response to precessional (23 and 19 k.y.) forcing (Figs. 8 and 11). Flood-ing surfaces (parasequence boundaries) inferred from the gamma log within the m5.34 sequence (Fig. 8) are ~25 k.y. in duration (i.e., 8 cycles in the 50 m of section shown on the inset rep-resenting <200 k.y.), consistent with precession forcing. If precessional forcing occurs, then it should be modulated by eccentricity forcing on the ~100 and 405 k.y. scale.

We suggest that although sequences may appear to be fractal and scale invariant (Schlager, 2004), they are in fact controlled by astronomi-cal forcing with distinct periodicities. Although we lack age control to unequivocally document 1.2 m.y., 405 k.y., or ~100 k.y. periodicities in our sequences, it is clear that glacioeustatic forcing occurred in the early to middle Miocene interval examined here (Fig. 12). Our chronol-ogy supports the existence of a 1.2 m.y. beat in early Miocene sequences and is consistent with a response on the 400 or 100 k.y. scale.

Paleodepth of Seismic Stratigraphic Features

Several issues remain to be addressed by Expedition 313 studies, including the infl uence of paleotopography of the clinothem on depo-sition (particularly lowstand deposits), paleo-relief between the clinoform infl ection and the bottomset, and the paleodepth of the rollovers and lowest point of onlap. Benthic foraminifera indicate that the bottomsets were deposited in ~100 m of water or slightly deeper. Sequences on the foresets are typically 150–200 m thick, with topsets as much as 200 ms (~200 m) above the bottomsets. This would imply greater water depth than indicated by benthic foraminifera. However, the role of loading on paleotopogra-phy (including two-dimensional effects) must be accounted for (Steckler et al., 1999). For example, two-dimensional backstripping shows that vertical differences in original geometry are muted compared to observed sediment thickness, especially in foresets (Kominz and Pekar, 2001).

We see no evidence for subaerial exposure on the clinothems sampled here (m5.8, m5.4 composite, and m5.2). Several sequences were sampled at the clinoform rollover: (1) m5.7 (which overlies m5.8) at Site M27, where the environments are coarsening-upward shoreface as part of a HST; (2) m5.33 at M28, where the environments are interpreted as shoreface coars-ening upward in the LST; and (3) m5.3 (which overlies m5.32) at M28, where the environ-ments are shoreface-offshore transition. Our observations are consistent with the recovery of lagoonal environments at ODP Site 1071 (Austin et al., 1998), 3 km landward of the m0.5 rollover. Together, this suggests that shorelines consistently move as far seaward as clinoform rollovers and that the depositional environment of the point of onlap at the clinoform rollover is nearshore in this area.

We sampled the lowest point of seismic onlap seaward of the rollover (refl ector 8) in sequence m5.34 at Site M28 (Figs. 6 and 7). Here, the onlap associated with the LST and

Page 20: Testing sequence stratigraphic models by drilling Miocene … · 2013-12-10 · Drilling foresets and testing sequence stratigraphic models Geosphere, October 2013 1237 confusion

Drilling foresets and testing sequence stratigraphic models

Geosphere, October 2013 1255

TS is a coarsening-upward offshore (50–75 m) environment. Sequence m5.33 was also sam-pled near the lowest point of onlap (between the sequence boundary and refl ector 13), where it consists of shoreface deposits (Figs. 6 and 7). These observations should prove to be useful in future work.

Back to Basics

Neal and Abreu (2009) eschewed the use of sea-level curves in recognizing systems tracts. Here we do not use relative sea-level curves in our interpretations of systems tracts; rather, we use basic seismic, core, and stratigraphic prin-ciples to recognize sequence boundaries, MFS, TS, and facies successions within sequences. We use facies successions and stratal surfaces to subdivide sequences into systems tracts. We focus on fi ning- and deepening-upward and coarsening- and shallowing-upward trends (Fig. 1) deciphered with lithologic and forami-niferal data that are applicable on topsets and foresets, but less applicable on bottomsets. Our simple predictive model of coarsening and fi n-ing trends (Fig. 1) is similar to the accommo-dation successions method of Neal and Abreu (2009) that focuses on progradational-aggra-dational-retrogradational patterns observed in seismic profi les (their Fig. 2). These com-plementary approaches allow objective rec-ognition of systems tracts that are not tied to preconceived notions.

CONCLUSIONS

We show that identification of seismic sequences using classic criteria is robust, allowing objective subdivision into sequences. Seismic sequence boundaries are recognized on topsets, foresets, and bottomsets and can be clearly differentiated from FSST and/or truncated HST and attendant surfaces. MFS can be generally inferred with seismic criteria as a downlap surface, although caution must be exercised in picking the stratigraphically lowest downlap surface as the MFS. We see little evidence for correlative conformities. Distinguishing LST and TST seismically is a challenging task. We show that interpreta-tion of systems tracts requires integration of seismic, core (lithology and foraminifera), and geophysical logs to develop unequivocal interpretations. Sequences embedded within million-year-scale composite sequences can be particularly challenging to interpret using seismic profi les alone. We note that our study area is consistent with preserving hierarchical orders of sequences on the tilt (1.2 m.y.) and eccentricity scales (100 and 405 k.y.).

ACKNOWLEDGMENTS

We thank the drillers and scientists of Integrated Ocean Drilling Program (IODP) Expedition 313 for their enthusiastic collaboration, the Bremen Core Repository for hosting our studies, and C. Lombardi, J. Criscione, and R. Miller for lithologic analyses. Funding was provided by COL/USSP, and samples were provided by the IODP and the International Continental Scientifi c Drilling Program. We thank B. Bracken and an anonymous reviewer for reviews, M. Kominz for discussions, and V. Abreu for bringing us back to basics.

REFERENCES CITED

Abdulah, K.C., and Anderson, J.B., 1994, Contrasting styles of deltaic deposition and the role of eustasy—Example from the Quaternary deltas of the Texas continental shelf, in Application of sequence stratigraphy to oil fi eld development: Hedberg Research Conference Pro-ceedings: Tulsa, Oklahoma, American Association of Petroleum Geologists, p. 1–11.

Aubry, M.-P., 1993, Neogene allostratigraphy and depo-sitional history of the DeSoto Canyon area, northern Gulf of Mexico: Micropaleontology, v. 39, p. 327–366, doi:10.2307/1485855.

Austin, J.A., Christie-Blick, N., Malone, M.J., and others, 1998, Initial Reports, Ocean Drilling Program, Leg 174A: College Station, Texas, Ocean Drilling Program, 324 p.

Boulila, S., Galbrun, B., Miller, K.G., Pekar, S.F., Brown-ing, J.V., Laskar, J., and Wright, J.D., 2011, On the ori-gin of Cenozoic and Mesozoic “third-order” eustatic sequences: Earth-Science Reviews, v. 109, p. 94–112, doi:10.1016/j.earscirev.2011.09.003.

Brown, L.F., and Fisher, W.L., 1977, Seismic-stratigraphic interpretation of depositional systems: Examples from Brazilian rift and pull-apart basins, in Payton, C.E., ed., Seismic stratigraphy—Applications to hydrocar-bon exploration: American Association of Petroleum Geologists Memoir 26, p. 213–248.

Browning, J.V., Miller, K.G., McLaughlin, P.P., Kominz, M.A., Sugarman, P.J., Monteverde, D., Feigenson, M.D., and Hernàndez, J.C., 2006, Quantifi cation of the effects of eustasy, subsidence, and sediment supply on Miocene sequences, Mid-Atlantic margin of the United States: Geological Society of America Bulletin, v. 118, p. 567–588, doi:10.1130/B25551.1.

Browning, J.V., Miller, K.G., Sugarman, P.J., Kominz, M.A., McLaughlin, P.P., and Kulpecz, A.A., 2008, 100 Myr record of sequences, sedimentary facies and sea-level change from Ocean Drilling Program onshore coreholes, U.S. Mid-Atlantic coastal plain: Basin Research, v. 20, p. 227–248, doi:10.1111/j.1365-2117.2008.00360.x.

Browning, J.V., Miller, K.G., Sugarman, P.J., Barron, J., McCarthy, F.M.G., Kulhanek, D.K., Katz, M.E., and Feigenson, M.D., 2013, Chronology of Eocene–Mio-cene sequences on the New Jersey shallow shelf: Implications for regional, interregional, and global cor-relations: Geosphere, doi:10.1130/GES00857.1.

Catuneanu, O., 2006, Principles of sequence stratigraphy: Elsevier, Amsterdam, 375 p.

Catuneanu, O., and 27 others, 2009, Towards the standardiza-tion of sequence stratigraphy: Earth-Science Reviews, v. 92, p. 1–33, doi:10.1016/j.earscirev.2008.10.003.

Christie-Blick, N., 1991, Onlap, offl ap, and the origin of unconformity-bounded depositional sequences: Marine Geology, v. 97, p. 35–56, doi:10.1016/0025-3227(91)90018-Y.

Christie-Blick, N., and Driscoll, N.W., 1995, Sequence stratigraphy: Annual Review of Earth and Planetary Sciences, v. 23, p. 451–478, doi:10.1146/annurev.ea.23.050195.002315.

Christie-Blick, N., Mountain, G.S., and Miller, K.G., 1988, Sea level history: Science, v. 241, p. 596, doi:10.1126/science.241.4865.596.

Christie-Blick, N., Mountain, G.S., and Miller, K.G., 1990, Seismic stratigraphic record of sea level change, in Geo-physics Study Committee, Sea-level change: National

Research Council Studies in Geophysics: Washington, D.C., National Academy Press, p. 116–140.

Coe, A.L., ed., 2003, The sedimentary record of sea level change: Cambridge, Cambridge University Press, 288 p.

Demarest, J.M., and Kraft, J.C., 1987, Stratigraphic record of Quaternary sea levels: Implications for more ancient strata, in Nummedal, D., et al., eds., Sea level fl uc-tuation and coastal evolution: Society of Economic Paleon tologists and Mineralogists Special Publication 41, p. 223–239, doi:10.2110/pec.87.41.0223.

DiCelma, C.N., Brunt, R.L., Hodgson, D.M., Flint, S.S., and Kavanagh, J.P., 2011, Spatial and temporal evolution of a Permian submarine slope channel-levee system, Karoo Basin, South Africa: Journal of Sedimentary Research, v. 81, p. 579–599, doi:10.2110/jsr.2011.49.

Embry, A.F., 2009, Practical sequence stratigraphy: Cana-dian Society of Petroleum Geologists, 81 p., www.cspg.org.

Flint, S.S., Hodgson, D.M., Sprague, A.R., Brunt, R.L., Van der Merwe, W.C., Figueiredo, J., Prélat, A., Box, D., Di Celma, C., and Kavanagh, J.P., 2011, Deposi-tional architecture and sequence stratigraphy of the Karoo basin fl oor to shelf edge succession, Laings-burg depocentre, South Africa: Marine and Petroleum Geology, v. 28, p. 658–674, doi:10.1016/j.marpetgeo.2010.06.008.

Galloway, W.E., 1989, Genetic stratigraphic sequences in basin analysis I: Architecture and genesis of fl ooding-surface bounded depositional units: American Associa-tion of Petroleum Geologists, Bulletin, v. 73, p. 125–142.

Gradstein, F., and 38 others, 2004, A geologic time scale 2004: Cambridge, Cambridge University Press, 589 p.

Greenlee, S.M., and Moore, T.C., 1988, Recognition and interpretation of depositional sequences and calculation of sea level changes from stratigraphic data—Offshore New Jersey and Alabama Tertiary, in Wilgus, C.K., et al., eds., Sea-level changes: An integrated approach: Society of Economic Paleontologists and Mineralo-gists Special Publication 42, p. 329–353, doi:10.2110/pec.88.01.0329.

Greenlee, S.M., Schroeder, F.W., and Vail, P.R., 1988, Seismic stratigraphic and geohistory analysis of Tertiary strata from the continental shelf off New Jersey—Calcula-tion of eustatic fl uctuations from stratigraphic data, in Sheridan, R., and Grow, J., eds., The Atlantic continental margin: U.S.: Boulder, Colorado, Geological Society of America, Geology of North America, v. I-2, p. 437–444.

Greenlee, S.M., Devlin, W.J., Miller, K.G., Mountain, G.S., and Flemings, P.B., 1992, Integrated sequence stratigra-phy of Neogene deposits, New Jersey continental shelf and slope: Comparison with the Exxon model: Geologi-cal Society of America Bulletin, v. 104, p. 1403–1411, doi:10.1130/0016-7606(1992)104<1403:ISSOND>2.3.CO;2.

Hancock, J.M., 1993, Transatlantic correlations in the Cam-panian-Maastrichtian stages by eustatic changes of sea-level, in Hailwood, E.A., and Kidd, R.B., eds., High resolution stratigraphy: Geological Society of London Special Publication 70, p. 241–256, doi:10.1144/GSL.SP.1993.070.01.17.

Hancock, J.M., and Kauffman, E.G., 1979, The great trans-gressions of the Late Cretaceous: Geological Society of London Journal, v. 136, p. 175–186, doi:10.1144/gsjgs.136.2.0175.

Hathaway, J.C., Poag, C.W., Valentine, P.C., Miller, R.E., Schultz. D.M., Manheim, R.T., Kohout, F.A., Bothner, M.H., and Sangrey, D.A., 1979, U.S. Geological Sur-vey core drilling on the Atlantic shelf: Science, v. 206, p. 515–527.

Heezen, B.C., Tharp, M., and Ewing, M., 1959, The fl oors of the oceans: I. The North Atlantic: Geological Soci-ety of America Special Paper 65, 122 p., doi:10.1130/SPE65-p1.

Holbourn, A.E., Kuhnt, W., Schulz, M., Flores, J.-A., and Andersen, N., 2007, Orbitally-paced climate evolution during the middle Miocene “Monterey” carbon-isotope excursion: Earth and Planetary Science Letters, v. 261, p. 534–550, doi:10.1016/j.epsl.2007.07.026.

Katz, M.E., Browning, J.V., Miller, K.G., Monteverde, D., Mountain, G.S., and Williams, R.H., 2013, Paleo-bathymetry and sequence stratigraphic interpretations 1 from benthic foraminifera: Insights on New Jersey

Page 21: Testing sequence stratigraphic models by drilling Miocene … · 2013-12-10 · Drilling foresets and testing sequence stratigraphic models Geosphere, October 2013 1237 confusion

Miller et al.

1256 Geosphere, October 2013

shelf architecture, IODP Expedition 313: Geosphere, doi:10.1130/GES00872.1.

Kidwell, S.M., 1989, Stratigraphic condensation of marine transgressive records: Origin of major shell deposits in the Miocene of Maryland: Journal of Geology, v. 97, p. 1–24, doi:10.1086/629278.

Kidwell, S.M., 1991, The stratigraphy of shell concentration, in Allison, P.A., and Briggs, D.E.G., eds., Taphon omy: Releasing the data locked in the fossil record: Topics in Geobiology Volume 9: New York, Springer-Verlag, p. 211–290.

Kominz, M.A., and Pekar, S.F., 2001, Oligocene eustasy from two-dimensional sequence stratigraphic backstrip-ping: Geological Society of America Bulletin, v. 113, p. 291–304, doi: 10.1130/0016-7606(2001)113<0291:OEFTDS>2.0.CO;2.

Loutit, T.S., Hardenbol, J., Vail, P.R., and Baum, G.R., 1988, Condensed section: The key to age determina-tion and correlation of continental margin sequences, in Wilgus, C.K., et al., eds., Sea-level changes: An inte-grated approach: Society of Economic Paleontologists and Mineralogists Special Publication 42, p. 183–213, doi:10.2110/pec.88.01.0183.

Miall, A.D., 1991, Stratigraphic sequences and their chrono-stratigraphic correlation: Journal of Sedimentary Petrology, v. 61, p. 497–505.

Miller, K.G., Mountain, G.S., Leg 150 Shipboard Party, and Members of the New Jersey Coastal Plain Drilling Project, 1996, Drilling and dating New Jersey Oligo-cene–Miocene sequences: Ice volume, global sea level, and Exxon records: Science, v. 271, p. 1092–1095, doi:10.1126/science.271.5252.1092.

Miller, K.G., Rufolo, S., Sugarman, P.J., Pekar, S.F., Brown-ing, J.V., and Gwynn, D.W., 1997, Early to middle Miocene sequences, systems tracts, and benthic forami-niferal biofacies, New Jersey coastal plain, in Miller, K.G., and Snyder, S.W., eds., Proceedings of the Ocean Drilling Program, Scientifi c Results, Leg 150X: College Station, Texas, Ocean Drilling Program, p. 169 186.

Miller, K.G., Mountain, G.S., Browning, J.V., Kominz, M., Sugarman, P.J., Christie-Blick, N., Katz, M.E., and Wright, J.D., 1998, Cenozoic global sea-level, sequences, and the New Jersey transect: Results from coastal plain and slope drilling: Reviews of Geophysics, v. 36, p. 569–601, doi:10.1029/98RG01624.

Miller, K.G., Browning, J.V., Mountain, G.S., Bassetti, M.A., Monteverde, D., Katz, M.E., Inwood, J., Lofi , J., and Proust, J.-N., 2013, Sequence boundaries are impedance contrasts: Core-seismic-log integration of Oligocene–Miocene sequences, New Jersey shallow shelf: Geosphere, v. 9, doi:10.1130/GES00858.1.

Mitchum, R.M., Jr., and Van Wagoner, J.C., 1991, High-frequency sequences and their stacking patterns; sequence-stratigraphic evidence of high-frequency eustatic cycles: Sedimentary Geology, v. 70, p. 131–160, doi:10.1016/0037-0738(91)90139-5.

Mitchum, R.M., Jr., Vail, P.R., and Thompson, S., III, 1977, Seismic stratigraphy and global changes of sea level; Part 2, The depositional sequence as a basic unit for stratigraphic analysis, in Payton, C.E., ed., Seismic stratigraphy; applications to hydrocarbon exploration: American Association of Petroleum Geologists Mem-oir 26, p. 53–62.

Monteverde, D.H., 2008, Sequence stratigraphic analysis of early and middle Miocene shelf progradation along the New Jersey margin [Ph.D. thesis]: New Brunswick, Rutgers University, 247 p.

Monteverde, D.H., Mountain, G.S., and Miller, K.G., 2008, Early Miocene sequence development across the New Jersey margin: Basin Research, v. 20, p. 249–267, doi:10.1111/j.1365-2117.2008.00351.x.

Mountain, G., and Monteverde, D., 2012, If you’ve got the time, we’ve got the depth: The importance of accurate

core-seismic correlation: American Geophysical Union Fall Meeting, abs. PP51B–2111.

Mountain, G.S., Miller, K.G., Blum, P., and Twitchell, D.C., et al., 1996, Proceedings of the Ocean Drilling Pro-gram, Scientifi c results, Volume 150: College Station, Texas, Ocean Drilling Program, 885 p.

Mountain, G.S., Burger, R.L., Delius, H., Fulthorpe, C.S., Austin, J.A., Goldberg, D.S., Steckler, M.S., McHugh, C.M., Miller, K.G., Monteverde, D.H., Orange, D.L., and Pratson, L.F., 2007, The long-term stratigraphic record on continental margins, in Nittrouer, C.A., et al., eds., Conti-nental margin sedimentation: From sediment transport to sequence stratigraphy: International Association of Sedi-mentologists Special Publication 37, p. 381–458.

Mountain, G.S., Proust, J.-N., McInroy, D., and Cotterill, C., and the Expedition 313 Scientists, 2010, Proceedings of the International Ocean Drilling Program, Expedi-tion 313: Tokyo, Integrated Ocean Drilling Program Management International, Inc., doi:10.2204/iodp.proc.313.2010.

Neal, J., and Abreu, V., 2009, Sequence stratigraphy hier-archy and the accommodation succession method: Geology, v. 37, p. 779–782, doi:10.1130/G25722A.1.

Olsson, R.K., Melillo, A.J., and Schreiber, B.L., 1987, Mio-cene sea level events in the Maryland coastal plain and the offshore Baltimore Canyon trough, in Ross, C., and Haman, D., eds., Timing and depositional history of eustatic sequences: Constraints on seismic stratig raphy: Cushman Foundation for Foraminiferal Research Spe-cial Publication 24, p. 85–97.

Owens, J.P., and Gohn, G.S., 1985, Depositional history of the Cretaceous series in the U.S. Atlantic coastal plain: Stratigraphy, paleoenvironments, and tectonic controls of sedimentation, in Poag, C.W., ed., Geologic evolu-tion of the United States Atlantic margin: New York, Van Nostrand Reinhold, p. 25–86.

Pälike, H., Frazier, J., and Zachos, J.C., 2006, Extended orbitally forced palaeoclimatic records from the equatorial Atlantic Ceara Rise: Quaternary Science Reviews, v. 25, p. 3138–3149, doi:10.1016/j.quascirev.2006.02.011.

Plint, A.G., and Nummedal, D., 2000, The falling stage sys-tems tract: Recognition and importance in sequence stratigraphic analysis, in Hunt, D., and Gawthorpe, R.L., eds., Sedimentary responses to forced regression: Geological Society of London Special Publication 172, p. 1–17, doi:10.1144/GSL.SP.2000.172.01.01.

Posamentier, H.W., and Vail, P.R., 1988, Eustatic controls on clastic deposition II—Sequence and systems tract models, in Wilgus, C.K., et al., eds., Sea level changes: An integrated approach: Society of Economic Paleon-tologists and Mineralogists Special Publication 42, p. 125–154, doi:10.2110/pec.88.01.0125.

Posamentier, H.W., Jervey, M.T., and Vail, P.R., 1988, Eustatic controls on clastic deposition I—Conceptual framework, in Wilgus, C.K., et al., eds., Sea level changes: An integrated approach: Society of Economic Paleontologists and Mineralogists Special Publication 42, p. 109–124, doi:10.2110/pec.88.01.0109.

Posamentier, H.W., Allen, G.P., James, D.P., and Tesson, M., 1992, Forced regressions in a sequence stratigraphic framework: Concepts, examples, and exploration sig-nifi cance: American Association of Petroleum Geolo-gists Bulletin, v. 76, p. 1687–1709.

Poulsen, C.J., Flemings, P.B., Robinson, R.A.J., and Metzger, J.M., 1998, Three-dimensional stratigraphic evolution of the Miocene Baltimore Canyon region: Implications for eustatic interpretations and the sys-tems tract model: Geological Society of America Bul-letin, v. 110, p. 1105–1122, doi:10.1130/0016-7606(1998)110<1105:TDSEOT>2.3.CO;2.

Proust, J.N., Mahieux, G., and Tessier, B., 2001, Field and seismic images of sharp-based shoreface deposits:

Implications for sequence stratigraphic analysis: Jour-nal of Sedimentary Research, v. 71, p. 944–957, doi:10.1306/041601710944.

Rabineau, M., Berné, S., Aslanian, D., Olivet, J.-L., Joseph, P., Guillocheau, F., Bourillet, J.-F., Ledrezen, E., and Granjeon, D., 2005, Sedimentary sequences in the Gulf of Lions: A record of 100,000 years climatic cycles: Marine and Petroleum Geology, v. 22, p. 775–804, doi:10.1016/j.marpetgeo.2005.03.010.

Schlager, W., 2004, Fractal nature of stratigraphic sequences: Geology, v. 32, p. 185–188, doi:10.1130/G20253.1.

Schlee, J.S., 1981, Seismic stratigraphy of Baltimore Can-yon Trough: American Association of Petroleum Geol-ogists Bulletin, v. 65, p. 26–53.

Sierro, F.J., and 14 others, 2009, Phase relationship between sea-level and abrupt climate change: Quaternary Science Reviews, v. 28, p. 2867–2881, doi:10.1016/j.quascirev.2009.07.019.

Snedden, J.W., and Liu, C., 2010, A compilation of Phanero-zoic sea-level change, coastal onlaps and recom-mended sequence designations: American Association of Petroleum Geologists Search and Discovery Article 40594, www.searchanddiscovery.com.

Steckler, M. S., Mountain, G. S., Miller, K.G., and Christie-Blick, N., 1999, Reconstruction of Tertiary prograda-tion and clinoform development on the New Jersey passive margin by 2-D backstripping: Marine Geology, v. 154, p. 399–420.

Sugarman, P.J., Miller, K.G., Owens, J.P., and Feigenson, M.D., 1993, Strontium-isotope and sequence stratigra-phy of the Miocene Kirkwood Formation, southern New Jersey: Geological Society of America Bulletin, v. 105, p. 423–436, doi:10.1130/0016-7606(1993)105<0423:SIASSO>2.3.CO;2.

Trincardi, F., and Correggiari, A., 2000, Quaternary forced-regression deposits in the Adriatic Basin and the record of composite sea-level cycles, in Hunt, D., and Gaw-thorpe, R., eds., Sedimentary responses to forced regres-sions: Geological Society of London Special Publication 172, p. 245–269, doi:10.1144/GSL.SP.2000.172.01.12.

Vail, P.R., 1987, Seismic stratigraphy interpretation using sequence stratigraphy: Part 1. Seismic stratigraphy interpretation procedure, in Bally A.W., ed., Atlas of seismic stratigraphy: American Association of Petro-leum Geologists Studies in Geology 27, p. 1–10.

Vail, P.R., Mitchum, R.M., Jr., and Thompson, S., III, 1977, Seismic stratigraphy and global changes of sea level; Part 4, Global cycles of relative changes of sea level, in Payton, C.E., ed., Seismic stratigraphy; applications to hydrocarbon exploration: American Association of Petroleum Geologists Memoir 26, p. 83–97.

Van Wagoner, J.C., Mitchum, R.M., Jr., Posamentier, H.W., and Vail, P.R., 1987, Seismic stratigraphy interpretation using sequence stratigraphy: Part 2, Key defi nitions of sequence stratigraphy, in Bally A.W., ed., Atlas of seis-mic stratigraphy: American Association of Petroleum Geologists Studies in Geology 27, p. 11–14.

Van Wagoner, J.C., Posamentier, H.W., Mitchum, R.M., Vail, P.R., Sarg, J.F., Loutit, T.S., and Hardenbol, J., 1988, An overview of the fundamentals of sequence stratig-raphy and key definitions, in Wilgus, C.K., et al., eds., Sea-level changes: An integrated approach: Society of Economic Paleontologists and Mineralogists Special Publication 42, p. 39–45, doi:10.2110/pec.88.01.0039.

Van Wagoner, J.C., Mitchum, R.M., Campion, K.M., and Rahmanian, V.D., 1990, Siliciclastic sequence stratig-raphy in well logs, cores, and outcrops: Concepts for high-resolution correlation of time and facies: Ameri-can Association of Petroleum Geologists Methods in Exploration Series 7, 55 p.

Wheeler, H.E., 1958, Time stratigraphy: American Associa-tion of Petroleum Geologists Bulletin, v. 42, p. 1047–1063.


Top Related