miocene relative sea level on the new jersey shallow...

20
Research Paper 1 Kominz et al. | Miocene relative sea level: IODP 313 New Jersey Margin GEOSPHERE | Volume 12 | Number 5 Miocene relative sea level on the New Jersey shallow continental shelf and coastal plain derived from one-dimensional backstripping: A case for both eustasy and epeirogeny M.A. Kominz 1 , K.G. Miller 2 , J.V. Browning 3 , M.E. Katz 4 , and G.S. Mountain 2 1 Department of Geosciences, Western Michigan University, 1186 Rood Hall, 1903 West Michigan Avenue, Kalamazoo, Michigan 49008, USA 2 Department of Earth and Planetary Sciences, and the Institute of Earth, Oceans, and Atmospheric Sciences, Rutgers, The State University of New Jersey, 610 Taylor Road, Piscataway, New Jersey 08854-8066, USA 3 Department of Earth and Planetary Sciences, Rutgers, The State University of New Jersey, 610 Taylor Road, Piscataway, New Jersey 08854-8066, USA 4 Department of Earth and Environmental Sciences, Rensselaer Polytechnic Institute, 110 8th Street, Troy, New York 12180, USA ABSTRACT Onshore drilling by Ocean Drilling Program (ODP) Legs 150X and 174AX and offshore drilling by Integrated Ocean Drilling Program (IODP) Expedition 313 provides continuous cores and logs of seismically imaged Lower to Middle Miocene sequences. We input ages and paleodepths of these sequences into one-dimensional backstripping equations, progressively accounting for the effects of compaction, Airy loading, and thermal subsidence. The resulting difference between observed subsidence and theoretical thermal subsidence provide relative sea-level curves that reflect both global average sea level and non-thermal subsidence. In contrast with expectations, backstripping sug- gests that the relative sea-level maxima in proximal onshore sites were lower than correlative maxima on the shelf. This requires that the onshore New Jer- sey coastal plain has subsided relative to the shelf, which is consistent with models of relative epeirogeny due to subduction of the Farallon plate. These models predict subsidence of the coastal plain relative to the shelf. Although onshore and offshore sea-level estimates are offset by epeirogeny, the ampli- tude of million-year–scale Early to Middle Miocene sea-level changes seen at the New Jersey margin is generally 5–20 m and occasionally as great as 50 m. These events are interpreted to represent eustatic variations, because they oc- cur on a shorter time frame than epeirogenic influences. Correction for epeiro- genic effects largely reconciles differences between onshore and offshore rela- tive sea-level estimates and suggests that backstripping provides a testable eustatic model for the Early to Middle Miocene. INTRODUCTION One of the outstanding challenges in studying Earth history is document- ing the timing and magnitude of global sea-level change (e.g., Haq et al., 1987, 1988; Miller et al., 2005). Here we use eustasy to mean the global change in sea level relative to a fixed point, e.g., the center of the Earth (Posamentier et al., 1988). We use the term relative sea-level change to include changes in eustasy coupled with epeirogenic (broad regional uplift; Grabau, 1936) and local changes in the height of lithosphere relative to the center of the Earth. That is, relative sea-level change is measured relative to a fixed point on the crust (Posamentier et al., 1988). Eustasy is of particular importance because it serves as the datum against which Earth’s tectonic and climate history can be measured (e.g., Miller et al., 2005). In this paper, we attempt to untangle thermal subsidence and epeirogeny from eustasy. The timing and magnitude of relative sea level has been studied in detail at the Late Cretaceous to Miocene of the New Jersey margin. Onshore New Jersey drilling by Ocean Drilling Program (ODP) Legs 150X and 174AX in con- cert with Legs 150 and 174A outer shelf and continental slope drilling (Fig. 1; e.g., Miller et al., 1997, 1998a; Mountain et al., 2010) has shown that the timing of sequence boundaries (erosional surfaces recognized in cores and seismic profiles), is consistent with d 18 O increases from deep-sea records (Miller et al., 1991, 1996a, 2005, 2011; Browning et al., 2008). This suggests that relative sea- level changes observed on the New Jersey margin were caused, at least in part, by eustatic variations due to ice growth and decay. Backstripping is a modeling technique that accounts for the effects of sedi- ment compaction, sediment loading, and, in this case, thermal subsidence. Applied to the onshore coreholes, backstripping provides a relative sea-level record for the New Jersey margin that in the absence of other tectonic effects yields a testable estimate of eustatic change (Miller et al., 2005; Kominz et al., 2008). However, mantle tomographic studies coupled with models of litho- spheric epeirogeny suggest that the New Jersey margin has undergone broad tectonic subsidence over the past 50 million years (e.g., Conrad et al., 2004; Moucha et al., 2008; Spasojević et al., 2008). Thus, more work is required to separate eustatic and epeirogenic effects in this region. Several processes must be accounted for to untangle eustasy from epeirog- eny. Most backstripping estimates of the magnitude of sea-level change in this region have used a one-dimensional approach that assumes an Airy response to sediment loads (Kominz et al., 1998; Miller et al., 1998a, 2005; Van Sickel GEOSPHERE GEOSPHERE; v. 12, no. 5 doi:10.1130/GES01241.1 8 figures; 3 tables; 5 supplemental files CORRESPONDENCE: [email protected] CITATION: Kominz, M.A., Miller, K.G., Browning, J.V., Katz, M.E., and Mountain, G.S., 2016, Miocene relative sea level on the New Jersey shallow conti- nental shelf and coastal plain derived from one- dimensional backstripping: A case for both eustasy and epeirogeny: Geosphere, v. 12, no. 5, p. 1– 20, doi:10.1130/GES01241.1. Received 31 July 2015 Revision received 9 May 2016 Accepted 3 August 2016 For permission to copy, contact Copyright Permissions, GSA, or [email protected]. © 2016 Geological Society of America THEMED ISSUE: Results of IODP Expedition 313: The History and Impact of Sea-Level Change Offshore New Jersey

Upload: others

Post on 21-Jun-2020

3 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Miocene relative sea level on the New Jersey shallow ...geology.rutgers.edu/images/stories/faculty/miller... · onshore and offshore sea-level estimates are offset by epeirogeny,

Research Paper

1Kominz et al. | Miocene relative sea level: IODP 313 New Jersey MarginGEOSPHERE | Volume 12 | Number 5

Miocene relative sea level on the New Jersey shallow continental shelf and coastal plain derived from one-dimensional backstripping: A case for both eustasy and epeirogenyM.A. Kominz1, K.G. Miller2, J.V. Browning3, M.E. Katz4, and G.S. Mountain2

1Department of Geosciences, Western Michigan University, 1186 Rood Hall, 1903 West Michigan Avenue, Kalamazoo, Michigan 49008, USA2 Department of Earth and Planetary Sciences, and the Institute of Earth, Oceans, and Atmospheric Sciences, Rutgers, The State University of New Jersey, 610 Taylor Road, Piscataway, New Jersey 08854-8066, USA

3Department of Earth and Planetary Sciences, Rutgers, The State University of New Jersey, 610 Taylor Road, Piscataway, New Jersey 08854-8066, USA4Department of Earth and Environmental Sciences, Rensselaer Polytechnic Institute, 110 8th Street, Troy, New York 12180, USA

ABSTRACT

Onshore drilling by Ocean Drilling Program (ODP) Legs 150X and 174AX and offshore drilling by Integrated Ocean Drilling Program (IODP) Expedition 313 provides continuous cores and logs of seismically imaged Lower to Middle Miocene sequences. We input ages and paleodepths of these sequences into one-dimensional backstripping equations, progressively accounting for the effects of compaction, Airy loading, and thermal subsidence. The resulting difference between observed subsidence and theoretical thermal subsidence provide relative sea-level curves that reflect both global average sea level and non-thermal subsidence. In contrast with expectations, backstripping sug-gests that the relative sea-level maxima in proximal onshore sites were lower than correlative maxima on the shelf. This requires that the onshore New Jer-sey coastal plain has subsided relative to the shelf, which is consistent with models of relative epeirogeny due to subduction of the Farallon plate. These models predict subsidence of the coastal plain relative to the shelf. Although onshore and offshore sea-level estimates are offset by epeirogeny, the ampli-tude of million-year–scale Early to Middle Miocene sea-level changes seen at the New Jersey margin is generally 5–20 m and occasionally as great as 50 m. These events are interpreted to represent eustatic variations, because they oc-cur on a shorter time frame than epeirogenic influences. Correction for epeiro-genic effects largely reconciles differences between onshore and offshore rela-tive sea-level estimates and suggests that backstripping provides a testable eustatic model for the Early to Middle Miocene.

INTRODUCTION

One of the outstanding challenges in studying Earth history is document-ing the timing and magnitude of global sea-level change (e.g., Haq et al., 1987, 1988; Miller et al., 2005). Here we use eustasy to mean the global change in sea level relative to a fixed point, e.g., the center of the Earth (Posamentier

et al., 1988). We use the term relative sea-level change to include changes in eustasy coupled with epeirogenic (broad regional uplift; Grabau, 1936) and local changes in the height of lithosphere relative to the center of the Earth. That is, relative sea-level change is measured relative to a fixed point on the crust (Posamentier et al., 1988). Eustasy is of particular importance because it serves as the datum against which Earth’s tectonic and climate history can be measured (e.g., Miller et al., 2005). In this paper, we attempt to untangle thermal subsidence and epeirogeny from eustasy.

The timing and magnitude of relative sea level has been studied in detail at the Late Cretaceous to Miocene of the New Jersey margin. Onshore New Jersey drilling by Ocean Drilling Program (ODP) Legs 150X and 174AX in con-cert with Legs 150 and 174A outer shelf and continental slope drilling (Fig. 1; e.g., Miller et al., 1997, 1998a; Mountain et al., 2010) has shown that the timing of sequence boundaries (erosional surfaces recognized in cores and seismic profiles), is consistent with d18O increases from deep-sea records (Miller et al., 1991, 1996a, 2005, 2011; Browning et al., 2008). This suggests that relative sea-level changes observed on the New Jersey margin were caused, at least in part, by eustatic variations due to ice growth and decay.

Backstripping is a modeling technique that accounts for the effects of sedi-ment compaction, sediment loading, and, in this case, thermal subsidence. Applied to the onshore coreholes, backstripping provides a relative sea-level record for the New Jersey margin that in the absence of other tectonic effects yields a testable estimate of eustatic change (Miller et al., 2005; Kominz et al., 2008). However, mantle tomographic studies coupled with models of litho-spheric epeirogeny suggest that the New Jersey margin has undergone broad tectonic subsidence over the past 50 million years (e.g., Conrad et al., 2004; Moucha et al., 2008; Spasojević et al., 2008). Thus, more work is required to separate eustatic and epeirogenic effects in this region.

Several processes must be accounted for to untangle eustasy from epeirog-eny. Most backstripping estimates of the magnitude of sea-level change in this region have used a one-dimensional approach that assumes an Airy response to sediment loads (Kominz et al., 1998; Miller et al., 1998a, 2005; Van Sickel

GEOSPHERE

GEOSPHERE; v. 12, no. 5

doi:10.1130/GES01241.1

8 figures; 3 tables; 5 supplemental files

CORRESPONDENCE: michelle .kominz@ wmich .edu

CITATION: Kominz, M.A., Miller, K.G., Browning, J.V., Katz, M.E., and Mountain, G.S., 2016, Miocene relative sea level on the New Jersey shallow conti­nental shelf and coastal plain derived from one­ dimensional backstripping: A case for both eustasy and epeirogeny: Geosphere, v. 12, no. 5, p. 1– 20, doi:10.1130/GES01241.1.

Received 31 July 2015Revision received 9 May 2016Accepted 3 August 2016

For permission to copy, contact Copyright Permissions, GSA, or [email protected].

© 2016 Geological Society of America

THEMED ISSUE: Results of IODP Expedition 313: The History and Impact of Sea-Level Change Offshore New Jersey

Page 2: Miocene relative sea level on the New Jersey shallow ...geology.rutgers.edu/images/stories/faculty/miller... · onshore and offshore sea-level estimates are offset by epeirogeny,

Research Paper

2Kominz et al. | Miocene relative sea level: IODP 313 New Jersey MarginGEOSPHERE | Volume 12 | Number 5

et al., 2004; Kominz et al., 2008). Only one model, of Upper Eocene to lower-most Miocene strata, used a two-dimensional backstripping approach that in-corporates flexural rigidity of the lithosphere to account for subsidence caused by sediment loads at a distance (Kominz and Pekar, 2001). There are several complications inherent in estimates of New Jersey margin relative sea level. One issue is the fact that coastal plain sediments rarely contain a complete record of sea-level change. They generally preserve only the transgressive and highstand systems tracts, leaving lowstand sediments farther seaward beneath what is now the continental shelf (Miller et al., 1998a). As discussed above, other tectonic effects have been postulated in this region so that tec-tonic subsidence may not be entirely thermal. In particular, the arrival of the Farallon slab beneath the North American east coast ca. 75 Ma requires that New Jersey subsided beyond the predicted thermal subsidence (Conrad et al., 2004; Kominz et al., 2008; Moucha et al., 2008; Müller et al., 2008; Spasojević et al., 2008). This means that the stratigraphic succession in this region has been imprinted by eustatic, thermal, and epeirogenic processes. Additionally, the whole Earth response to loading effects of water coupled with the unload-

ing of glaciers (glacial isostatic adjustment [GIA]) has been shown to vary globally (e.g., Peltier, 1998) with impact in this region (e.g., Raymo et al., 2011). This effect is most pronounced during the large Northern Hemisphere ice ages of the past 2.7 m.y., but GIA influences the reference frame of older records as well (e.g., Raymo et al., 2011).

Drilling data provided by IODP Expedition 313 (hereafter “Exp 313”) on the New Jersey shallow continental shelf focused on Lower to Middle Mio-cene strata (Mountain et al., 2010). This data set presents an opportunity to estimate amplitudes of offshore relative sea-level change based on cores, logs, and seismic profiles that can be compared to the onshore results. By providing estimates of the magnitude of relative sea-level change, we can begin to address the magnitude and timing of both eustasy and more regional epeirogeny.

Exp 313 drilled three coreholes (sites M27, M28, and M29; Fig. 1) in ~30 m of water, targeting Miocene sequences that also were cored in multiple loca-tions on the onshore coastal plain. The Miocene section is well imaged on a grid of seismic profiles that show a series of clinothems, which are packages

100

200

1000

2000

3000

C O N T I N E N T A L S H E L F

NewJersey

aula

PA L E O Z O I C OU T C R O P

C R E TA C E O U S

MI O C E N E

M27M28

M29

Atlantic

Cape

Cape

Bass RiverAncora

Bethany Beach

Millville

Ft. Mott 1071

1072 1073

Beach

View

Island

SeaGirt

OceanCity

903 904

905

902

906MayZoo

May

Drillsites

onshore ODP

AMCORDSDP

oil exploration

IODP Expedition 313

Seismic Profiles

Figure 2Drill Sites

offshore ODP

*DickinsonAnchor

wellCH0698 Ew9009

Oc270

N

39°

40°

–73° –72°–75° –74°

Figure 1. Location map. Sites used in this work are indicated as large red circles for Integrated Ocean Drilling Program (IODP) Expedition 313 sites and as large green cir-cles for the onshore coreholes with lower to middle Miocene strata. Seismic profiles are from three different data acquisition cruises (R/V Ewing cruise Ew9009, R/V Oceanus cruise Oc270, and R/V Cape Hatteras cruise CH0698; Monteverde et al., 2008; Mountain et al., 2010; Miller et al., 2013a). The seismic section Oc270 line 529 (red line passing through Exp 313 drill sites) is shown in Figure 2. The Anchor Dickinson well (gray filled dot; Poag, 1985; Sugarman et al., 2011) was drilled as a gas exploration well between the Cape May and Cape May Zoo sites. AMCOR—Atlan-tic Margin Coring Project; DSDP—Deep Sea Drilling Program.

Page 3: Miocene relative sea level on the New Jersey shallow ...geology.rutgers.edu/images/stories/faculty/miller... · onshore and offshore sea-level estimates are offset by epeirogeny,

Research Paper

3Kominz et al. | Miocene relative sea level: IODP 313 New Jersey MarginGEOSPHERE | Volume 12 | Number 5

of sediment that prograde seaward and are bounded by surfaces (in this case, sequence boundaries) with distinct sigmoidal (clinoform) geometries (Fig. 2; Mountain et al., 2010). These sites, drilled as a transect along seismic line Oc270 529, provide a two-dimensional cross section of several stratigraphic sequences formed between 12 Ma and 22 Ma (Mountain et al., 2010). Here, we present the results of a one-dimensional backstripping study of this new data set. As such, it is directly comparable to the one-dimensional modeling already published from the coastal plain (Kominz et al., 2008). By extending that study to this offshore location, it is possible to provide a preliminary esti-mate of the magnitude of million-year–scale relative sea-level changes and to consider the effects of the Farallon slab and GIA on relative sea-level change across the New Jersey margin.

METHODS AND INPUT DATA

Backstripping utilizes compaction, age, and paleodepth observations from cores or outcrops to estimate how the basement would have subsided (tec-tonic subsidence [TS]) in the absence of sediments and eustatic sea-level change (DSL) (Steckler and Watts, 1978; Bond and Kominz, 1984):

TS = Φ S*ρm − ρs

ρm − ρw− ∆SL

ρw

ρm − ρw− ∆SL +WD

, (1)

where: r is density of the asthenospheric mantle (m), the decompacted sedi-ment (s), and seawater (w); WD is local water depth; S* is the decompacted sediment thickness; and F is the basement response function.

IODP Expedition 313

10000 1000 cdp20003000400050006000700080009000

0

.1

.2

.3

.4

.5

.6

.7

.8

.9

1.0 1.0

0

Oc270 529

.1

.2

.3

.4

.5

.6

.7

.8

.9

Sec

onds

0 10 km

NW SE27 28 29

m4

m4.1

m4.1

m4.5

m4.2m4.3

m4.4

m5

m5

m5.2

m5.2

m5.4

m5.4

m5.6

m5.6

m5.47

m5.45

m5.7

m5.8m6

K/To1

o.5

m5.3

m5.3m5.32

m5.33

m5.34

m1

m3 Figure 2. Seismic profile Oc270 529. Seis-mic interpretations from Monteverde et al. (2008) and Mountain et al. (2010). Deposi-tional sequences are highlighted in vari-ous colors and named according to the underlying sequence boundary. For exam-ple, at M27 the strata colored blue rest on sequence boundary m5.4 and belong to sequence m5.4. IODP—Integrated Ocean Drilling Program.

Page 4: Miocene relative sea level on the New Jersey shallow ...geology.rutgers.edu/images/stories/faculty/miller... · onshore and offshore sea-level estimates are offset by epeirogeny,

Research Paper

4Kominz et al. | Miocene relative sea level: IODP 313 New Jersey MarginGEOSPHERE | Volume 12 | Number 5

In practice, backstripping begins with measuring porosity as a function of depth and lithology in a borehole or outcrop. These data are used to estimate the porosity, and thus, decompacted thickness (S*) of every part of the sedi-mentary column through time. The sediment column is removed at each time step and replaced by a column of seawater that represents the hole that sedi-ment of the estimated porosity, thickness, and density would have filled. In calculating the effect of the sediment load, we have applied a one-dimensional approach that assumes the underlying lithosphere has no rigidity (the Airy model, in which F, in Equation 1 is taken as 1). We also account for changes in water depth as described below. For that, paleodepth estimates (WD) are added to the water column that was filled with sediment (above) to determine the total subsidence of the basement in the absence of sediment. The result is termed the first reduction, or R1 (Bond et al., 1989):

R1= S*ρm − ρs

ρm − ρw+ WD

, (2)

where the variables are the same as described in Equation 1. R1 does not take into account global average sea-level change ( = eustasy; DSL in Equation 1). R1 is based on observed data that can be obtained from a corehole with a reasonable degree of accuracy.

Tectonism at a passive margin may be assumed to follow the theoretical behavior of a cooling plate due to lithospheric stretching during rifting, as demonstrated by the fact that long-term (10–100 m.y.) passive margin records can be largely explained with an exponential fit (e.g., McKenzie, 1978; Royden and Keen, 1980). Because of the predictable nature of tectonics in this setting, we can estimate eustasy by separating the long-term, thermal component of subsidence from any perturbations observed in the R1 curve. A second param-eter, R2, is calculated as the difference between R1 and a cooling plate model fit to the observed R1 data. We fit a thermal plate model (following Kominz et al., 2008) to the R1 curve to determine the thermal component of subsid-ence. The difference between the estimated thermal subsidence (taken as TS, Equation 1) and R1 (Equation 2), after being corrected for the change in water load (again assuming a one-dimensional Airy response of the underlying litho-sphere) yields the second reduction R2:

R2 = (R1−TS) ρw

ρm − ρw

. (3)

Employing Equations 1 and 2 to define TS and R1 explicitly defines R2 as a eustatic estimate. However, this is only true if all tectonics at the bore-hole in question are governed by thermal subsidence. Additional errors may arise as a result of assuming Airy isostasy, in a basin that actually responded flexurally to the sediment load because subsidence is under-esti mated at the locus of loading and overestimated at the periphery of the basin. However, in our approach we are looking at the difference between R1 and the theoretical thermal subsidence fit to the observed subsidence, R1. The form of subsidence resulting from the flexural response to loading of a thermal basin is thermal in both the center of the basin and its periph-

ery (Bond et al., 1988). Therefore, the difference between R1 and thermal tectonics, R2, is not affected by one-dimensional backstripping (Kominz et al., 1998). What is lost through one-dimensional analysis of a flexurally subsiding thermal basin is the detailed relationship between data points along a two-dimensional profile that could add insights for interpretations (e.g., Steckler et al., 1999; Kominz and Pekar, 2001). Observed relative sea level is also complicated by glacial isostatic adjustment (GIA; Peltier, 1998), by gravitational, rotational, and flexural effects due to changing ice sheets (collectively known as “static equilibrium” effects; Kopp, et al., 2010) and by dynamic topography (e.g., Milne et al., 2009). These issues will be consid-ered in evaluating the results.

Lithology and Porosity

We use the backstripping approach of Bond et al. (1989) in which poros-ity is assumed to be lithology dependent and the porosity of mixed lithol-ogies is calculated based on the proportion of lithologies present. This has been shown to be a viable approach for marine sediments (Kominz et al., 2011). Porosities based on moisture and density measurements from the three sites used in this study were combined with smear slide inter-pretations, grain-size analysis of discrete samples, and downhole log data (Mountain, et al., 2010; Miller et al., 2013a; Ando et al., 2014) to obtain lithol-ogy-dependent porosity versus depth plots for the New Jersey shelf (Fig. 3 and Table 1). Lithologies were extended from the depths of each discrete measurement to the nearest stratigraphic boundary (Fig. 2) above and be-low that measurement; otherwise, the lithologic boundaries were placed between samples. Siliciclastic sand-, silt- and clay-dominated intervals are sufficiently distinct in the Exp 313 cores to provide valid comparisons with standard, lithology-based porosity versus depth plots published elsewhere. While the Exp 313 data show no consistent pattern with depth, their trends are close to the >90% clay, the >90% silt, and the >60% sand curves based on a global Ocean Drilling Program (ODP) database (Kominz et al., 2011). Thus, we use the Kominz et al. (2011) porosity versus depth relations to decompact the sediments in these coreholes (Fig. 3). High-end and low-end porosity versus depth curves that encompass the range of observed porosities were also applied for comparison (see Supplemental Table S11 and Supplemental Fig. S12).

Ages and Environments of Deposition

The focus of this work is the Lower–Middle Miocene section, for which considerable data are available and has been synthesized to provide input for modeling (Table 2; detailed input data are provided in Supplemental Table S23). Age estimates are based on Browning et al. (2013), who used calcare-ous nannofossil, dinoflagellate cyst, and diatom biostratigraphy, and numer-

Table S-1. Porosity vs. Depth Curves

Clay best Porosity = 82.7 e-depth/418 < 172 mClay best Porosity = 61.4 e-depth/1676 > 172 mClay low Porosity = 45 e-depth/417 < 172 mClay low Porosity = 33 e-depth/1674 > 172 mClay high Porosity = 90 e-depth/425 < 95Clay high Porosity = 76 e-depth/1674 > 95

Silt best Porosity = 75 e-depth/419 < 97mSilt best Porosity = 63 e-depth/1675 > 97mSilt low Porosity = 60 e-depth/420 < 80Silt low Porosity = 52 e-depth/1676 > 80Silt high Porosity = 85 e-depth/418 < 63Silt high Porosity = 76 e-depth/1677 > 63

Sand best Porosity = 54.5 e-depth/1648

Sand low Porosity = 40 e-depth/2164

Sand high Porosity = 70 e-depth/3566

Carbonate best Porosity = 88 e-depth/1338

Carbonate low Porosity = 25 e-depth/3126

Carbonate high Porosity = 90 e-depth/1730

6050403020 70 800

600

500

400

300

200

100

700

800

Porosity Data

Clay

Silt

Sandstone

Porosity Relations

Clay

Silt

Sandstone

Porosity (%)

Dep

th (m

)

0 605040302010 70 80

0 605040302010 70 80 90

TABLE S2C: INPUT DATA M29Thickness age Density Clay Micrite Sand CaCO 3 Silt WD low WD mid WD high

PH 3.535 0.010 2.705 0.235 0.000 0.060 0.000 0.705 36.000 36.000 36.000P 3.590 0.070 2.705 0.235 0.000 0.060 0.000 0.705 10.000 25.000 40.000Pleist 2 1.755 2.691 0.185 0.000 0.260 0.000 0.555 10.000 25.000 40.000Pleist 3 2.240 2.661 0.098 0.000 0.610 0.000 0.293 10.000 25.000 40.000Pleist 4 0.545 2.707 0.243 0.000 0.030 0.000 0.728 10.000 25.000 40.000Pleist 5 0.820 2.710 0.250 0.000 0.000 0.000 0.750 10.000 25.000 40.000Pleist 6 0.915 2.703 0.235 0.000 0.060 0.000 0.705 10.000 25.000 40.000Pleist 7 0.620 2.657 0.108 0.000 0.520 0.050 0.323 10.000 25.000 40.000Pleist 8 1.480 2.632 0.095 0.000 0.570 0.050 0.285 10.000 25.000 40.000P1 0.000 0.090 2.700 0.000 0.000 0.000 0.000 0.000 10.000 25.000 40.000P1 0.135 0.100 2.632 0.095 0.000 0.570 0.050 0.285 5.000 10.000 20.000Pleist 9 13.005 2.611 0.008 0.000 0.960 0.010 0.023 5.000 10.000 20.000Pleist 10 14.360 2.613 0.013 0.000 0.950 0.000 0.038 5.000 10.000 20.000P2 0.000 0.110 2.700 0.000 0.000 0.000 0.000 0.000 5.000 10.000 20.000P2 2.105 1.030 2.613 0.013 0.000 0.950 0.000 0.038 5.000 10.000 20.000Pleist 11 4.895 2.645 0.038 0.000 0.850 0.000 0.113 5.000 10.000 20.000UlP 0.000 1.080 2.700 0.000 0.000 0.000 0.000 0.000 5.000 10.000 20.000

1Supplemental Table S1. Porosity versus depth curves. Please visit http:// dx .doi .org /10 .1130 /GES01241 .S1 or the full-text article on www .gsapubs .org to view Sup-plemental Table S1.

2Supplemental Figure S1. Integrated Ocean Drilling Program Expedition 313 porosity versus depth for sediments dominated by (>50%) sands, silts, and clay. Please visit http:// dx .doi .org /10 .1130 /GES01241 .S2 or the full-text article on www .gsapubs .org to view Supplemental Figure S1.

3Supplemental Table S2. Ages, densities (rho), lithol-ogies, and water depths used as input for sequences from coreholes M27 (section A), M28 (section B), and M29 (section C). Please visit http:// dx .doi .org /10 .1130 /GES01241 .S3 or the full-text article on www .gsapubs .org to view Supplemental Table S2.

Page 5: Miocene relative sea level on the New Jersey shallow ...geology.rutgers.edu/images/stories/faculty/miller... · onshore and offshore sea-level estimates are offset by epeirogeny,

Research Paper

5Kominz et al. | Miocene relative sea level: IODP 313 New Jersey MarginGEOSPHERE | Volume 12 | Number 5

ous strontium (Sr) isotopic age estimates of calcium carbonate from mollusk shells, shell fragments, and foraminiferal tests. Ages were assigned using the time scale of Gradstein et al. (2012; GTS2012). By using Sr-isotope stra-tigraphy, Browning et al. (2013) overcame the challenges of poor magneto-stratigraphy and poor biostratigraphy posed by coarse, clastic, nearshore sediments. They were able to generate age resolution of typically ±0.5 m.y., and in many sequences, age resolution was as good as ±0.25 m.y., where age resolution refers to uncertainties in correlation to the geologic time scale (Browning et al., 2013).

Superposition and sequence stratigraphic principles provide additional relative age constraints. Our biostratigraphic and Sr-isotopic data have been combined with seismic profiles, downhole logs, sedimentological data, and sequence stratigraphic interpretations to provide age estimates for sequence boundaries and several prominent subsequence stratal surfaces (Browning et al., 2013). In this study, dates of some sequence boundaries and maximum flooding surfaces (MFS) have been shifted slightly from previous work of

Mountain et al. (2010) and Browning et al. (2013) to honor new, detailed se-quence stratigraphic correlations by Miller et al. (2013b) and to reconcile minor differences between onshore and offshore locations (Table 3). In particular, sequence boundaries are assumed to represent times of non-deposition that may represent more time at one location than another, but the sediments be-low the unconformity are everywhere older than those above. Additionally, MFS (based on core data) are assumed to be correlative and are given a spe-cific date within the sequence. Finally, where onshore and offshore sequences are correlative, minor shifts in ages (0–0.45 m.y.) of the offshore strata were made to align correlative sequences. All assigned dates are within the error limits presented by Browning et al. (2013).

Environments of deposition at the three well sites were based on benthic foraminiferal assemblages, planktonic foraminiferal abundances, and palyno-logical proxies (Katz et al., 2013; McCarthy et al., 2013; Miller et al., 2013b) coupled with lithofacies interpretations (Browning et al., 2013). Generally the lithofacies environments are: foreshore (<10 m), upper shoreface (10–20 m), shoreface-offshore transition (20–30 m), and offshore (below storm wave base that we adopt as ≥30 m). Benthic foraminiferal bathymetric zones are defined as inner neritic (0–30 m), middle neritic (30–100 m), and outer neritic (100–200 m) (van Morkhoven et al., 1986). Miller et al. (1997) and Katz et al. (2013) used a subdivision of these zones on the New Jersey margin when they in-terpreted Elphidium-dominated biofacies as <10 m, Hanzawaia hughesi-domi-nated biofacies as 10–25 m, Pseudononion pizarrensis–dominated biofacies as 25–50 m, Bulimina gracilis–dominated biofacies as 50–80 m, and Uvigerina- dominated biofacies as >75 m. For deeper biofacies not recovered by Miller et al. (1997), Katz et al. (2013) used key taxa (e.g., Cibicidoides pachyderma, Cibicidoides primulus, Hanzawaia mantaensis, and Oridorsalis umbonatus) often found in high-diversity, low-dominance assemblages that indicate outer neritic paleodepths (100–200 m) (e.g., Parker, 1948; Poag 1981; van Morkhoven et al., 1986; Katz et al., 2003). Because changes from the Uvigerina spp. bio-facies to the deeper biofacies occur gradually within sequences, Katz et al. (2013) inferred that the maximum water depths are in the shallower part of the outer neritic zone (~120 m). Water-depth ranges as well as best estimate water depths are used in modeling.

0 605040302010 70 80 900

600

500

400

300

200

100

700

800

Porosity Data

Clay

Silt

Sandstone

Porosity Relations

Clay

Silt

Sandstone

Porosity (%)

Dep

th (m

)

Figure 3. Porosity versus depth measured in Integrated Ocean Drilling Pro-gram (IODP) Expedition 313 sediments composed of >50% sand (yellow circles), >50% silt (green diamonds), or >50% clay-sized particles (brown squares.) The yellow, green, and brown lines show porosity-depth rela-tions from a database of measurements in other ODP boreholes (Kominz et al., 2011) used for decompacting each sediment type in this study. See text for discussion and Table 1 for curves used in modeling. Porosity ver-sus depth curves were also generated to evaluate the impact of the full range of porosity data on results and are provided in Supplemental Table S1 and Supplemental Figure S1 (footnotes 1 and 2, respectively).

TABLE 1. POROSITY VERSUS DEPTH CURVES

Clay Porosity = 82.7 e–depth/418 <172 mClay Porosity = 61.4 e–depth/1676 >172 mSilt Porosity = 75 e–depth/419 <97 mSilt Porosity = 63 e–depth/1675 >97 mSand Porosity = 54.5 e–depth/1648

Carbonate Porosity = 88 e–depth/1338

Note: The porosity versus depth curves used in this model are provided above. Porosity values, e.g., porosity at the surface and calculated porosity, are in volume percent. Depths, decay constant, and depth ranges are given in meters (m). Porosity curves defined to encompass the full range of data are provided in Supplemental Table S3 (see footnote 1) and plotted in Supplemental Fig. S1 (see footnote 2).

Page 6: Miocene relative sea level on the New Jersey shallow ...geology.rutgers.edu/images/stories/faculty/miller... · onshore and offshore sea-level estimates are offset by epeirogeny,

Research Paper

6Kominz et al. | Miocene relative sea level: IODP 313 New Jersey MarginGEOSPHERE | Volume 12 | Number 5

TABLE 2. AGES, AVERAGE DENSITIES (RHO), AVERAGE LITHOLOGIES, AND WATER DEPTHS* USED AS INPUT FOR SEQUENCES FROM SITES M27, M28, AND M29

Depth Age RhoAverage

clayAverage

sandAverageCsand

Averagesilt WD low WD mid WD high

M27 input

holo 0 0.01 2.66 0.07 0.79 0.02 0.1 34 34 34P 0.28 0.011 2.66 0.07 0.79 0.02 0.1 34 34 34P 0.28 0.072 0 5 10P 13.57 0.085 2.66 0.13 0.65 0.02 0.19 0 5 10P 13.57 0.19 20 35 55P 22.8 0.22 2.68 0.26 0.33 0.01 0.39 20 35 55P 22.8 1.03 0 5 10P 27.41 1.08 2.66 0.14 0.62 0.02 0.21 0 5 10P 27.41 1.4 0 5 10MP 32.93 1.5 2.66 0.13 0.65 0.01 0.2 0 5 10MP 32.93 11.4 –5 0 5m1 97.03 11.8 2.66 0.13 0.67 0 0.19 –5 0 5m1 97.03 12 –5 0 5m3 112.07 12.3 2.67 0.34 0.15 0 0.51 –5 0 5m3 112.07 12.6 –5 0 5m4 136.03 12.7 2.68 0.21 0.48 0 0.31 –5 0 5m4.1 136.03 12.96 –5 0 5m4.2 180.71 13 2.67 0.18 0.55 0 0.27 5 10 20m4.2 184.09 13.1 15 30 50m4.3 209 13.2 2.69 0.28 0.23 0 0.48 20 35 55m4.4 209 13.4 15 30 50m4.5 218.39 13.5 2.69 0.1 0.33 0.01 0.55 5 10 20m4.5 218.39 13.65 20 35 55m5 225 13.75 2.66 0.1 0.56 0.01 0.33 20 35 55m5 225 14.8 10 20 35onlap uncf 227.51 14.85 2.69 0.11 0.34 0.06 0.47 35.1 65.9 97.6downlap uncf 227.51 14.9 35.1 65.9 97.6MFS 228 15.2 2.7 0.09 0.14 0.01 0.76 40 75 110m5.2 236.15 15.3 2.67 0.06 0.46 0 0.47 5 10 20m5.2 236.15 15.6 20 35 55MFS 249.76 15.8 2.7 0.09 0.08 0.02 0.81 25 75 80m5.3 256.31 15.95 2.7 0.09 0.15 0.02 0.75 15 30 50m5.3 256.31 16.65 35 45 55MFS 264.61 16.7 2.7 0.09 0.12 0.01 0.78 35 45 55m5.33 271.35 16.75 2.69 0.07 0.34 0.01 0.58 25 30 50m5.33 271.35 17.1 10 15 25MFS 288.64 17.3 2.69 0.09 0.38 0.01 0.52 35 40 50m5.34 295.12 17.35 2.69 0.17 0.3 0.02 0.52 35 40 50m5.34 295.12 17.65 10 25 40MFS 297.7 17.67 2.7 0.14 0.1 0 0.76 0 50 80m5.4 314.67 17.7 2.7 0.11 0.15 0.01 0.73 10 25 40m5.4 315.78 17.79 20 40 70m5.45 336.17 17.85 2.69 0.09 0.29 0.02 0.6 25 40 60m5.45 336.17 17.86 5 10 20m5.47 355.64 17.9 2.64 0.05 0.83 0 0.12 5 10 20

(continued )

Page 7: Miocene relative sea level on the New Jersey shallow ...geology.rutgers.edu/images/stories/faculty/miller... · onshore and offshore sea-level estimates are offset by epeirogeny,

Research Paper

7Kominz et al. | Miocene relative sea level: IODP 313 New Jersey MarginGEOSPHERE | Volume 12 | Number 5

TABLE 2. AGES, AVERAGE DENSITIES (RHO), AVERAGE LITHOLOGIES, AND WATER DEPTHS* USED AS INPUT FOR SEQUENCES FROM SITES M27, M28, AND M29 (continued )

Depth Age RhoAverage

clayAverage

sandAverageCsand

Averagesilt WD low WD mid WD high

M27 input (continued )

m5.6 355.64 18.6 0 5 15m5.7 361.39 18.8 2.67 0.13 0.57 0 0.3 5 10 20m5.7 361.39 19.1 0 5 10MFS 434.96 19.5 2.67 0.4 0.59 0 0 40 75 110TransSurf 477.79 19.7 2.69 0.86 0.13 0 0 20 35 55m5.8 494.98 20.1 2.67 0.28 0.71 0 0 30 50 80m5.8 494.98 20.7 40 75 110m6 515.11 20.9 2.68 0.13 0.53 0.02 0.31 40 75 110m6 515.11 23 40 75 110o6 538.79 23.5 2.67 0.13 0.55 0.02 0.29 40 75 110o6 538.79 28.2 40 75 110o3 617.11 29.3 2.68 0.17 0.4 0.02 0.41 40 75 110o3 617.11 32.2 40 75 110o1 625.94 32.3 2.7 0.21 0.3 0 0.49 40 75 110o1 625.94 33.6 30 50 80ba 629.44 33.8 2.71 0.26 0.03 0.01 0.61 30 50 80eo 629.44 55 50 70 100

M28 input

holo 0 0.01 2.66 0.02 0.9 0 0.08 35 35 35Mio 223.7 10 2.7 0.2 0.01 0 0.79 25 40 60m4.2 223.7 13.1 30 50 80m4.3 244 13.2 2.7 0.2 0.02 0 0.79 30 50 80m4.4 244 13.36 0 5 10MFS 250.6 13.4 2.65 0.09 0.53 0.01 0.36 20 35 55m4.5 254.2 13.55 2.7 0.18 0.08 0 0.73 10 25 40m4.5 254.2 13.56 0 5 10m5 276.8 13.7 2.66 0.03 0.84 0.02 0.12 10 25 40m5 276.8 14.65 5 10 25onlap uncf 303.6 14.85 2.63 0.08 0.56 0.01 0.35 19.3 36.6 54downlap uncf 303.6 14.9 19.3 36.6 54MFS 310.2 15.2 2.71 0.1 0.09 0 0.81 30 50 70m5.2 323.2 15.4 2.69 0.12 0.31 0 0.56 10 25 30m5.2 323.2 15.7 10 15 25MFS 330.4 15.8 2.55 0.02 0.87 0.01 0.1 10 25 40m5.3 361 16 2.64 0.02 0.91 0.01 0.07 10 25 40m5.3 361 16.5 5 10 20downlap uncf 386.2 16.55 2.66 0.02 0.91 0 0.07 9.1 22.4 36.6downlap uncf 386.2 16.6 9.1 22.4 36.6MFS 391 16.7 2.66 0.03 0.85 0 0.11 10 25 40m5.33 404 16.8 2.65 0.02 0.9 0.01 0.06 5 10 20m5.33 404 17.05 10 25 40MFS 449 17.3 2.65 0.06 0.66 0 0.27 30 50 80m5.34 479 17.45 2.65 0.07 0.46 0 0.47 30 50 80m5.34 479 17.6 40 75 110

(continued )

Page 8: Miocene relative sea level on the New Jersey shallow ...geology.rutgers.edu/images/stories/faculty/miller... · onshore and offshore sea-level estimates are offset by epeirogeny,

Research Paper

8Kominz et al. | Miocene relative sea level: IODP 313 New Jersey MarginGEOSPHERE | Volume 12 | Number 5

TABLE 2. AGES, AVERAGE DENSITIES (RHO), AVERAGE LITHOLOGIES, AND WATER DEPTHS* USED AS INPUT FOR SEQUENCES FROM SITES M27, M28, AND M29 (continued )

Depth Age RhoAverage

clayAverage

sandAverageCsand

Averagesilt WD low WD mid WD high

M28 input (continued )

MFS 504.9 17.67 2.68 0.08 0.35 0 0.57 40 75 110m5.4 512.3 17.7 2.65 0.03 0.71 0 0.26 30 50 80m5.4 512.3 17.79 40 75 110m5.45 533.6 17.85 2.65 0.07 0.58 0 0.35 40 75 110m5.45 533.6 17.9 40 75 110m5.47 545.5 18 2.65 0.05 0.83 0 0.12 40 75 110m5.47 545.5 18.1 40 75 110m5.6 567.5 18.3 2.66 0.04 0.87 0 0.09 40 75 110m5.6 567.5 18.85 40 75 110m5.7 611.6 19 2.66 0.06 0.81 0 0.13 40 75 110m5.7 611.6 19.4 10 25 40MFS 654.3 19.5 2.71 0.87 0.13 0 0 45 50 75m5.8 663 19.7 2.66 0.53 0.47 0 0 45 50 75m5.8 663 20.5 30 50 80m6 668.6 20.9 2.65 0.23 0.23 0 0.54 30 50 80

M29 input

PH 0 0.01 2.71 0.24 0.06 0 0.71 36 36 36P 3.54 0.07 10 25 40P1 15.5 0.09 2.68 0.17 0.29 0.01 0.52 10 25 40P1 15.5 0.1 5 10 20P2 43 0.11 2.62 0.04 0.83 0.02 0.12 5 10 20P2 43 1.03 5 10 20UlP 50 1.08 2.63 0.03 0.9 0 0.08 5 10 20UlP 50 11.4 –10 –5 0m1 160 11.8 2.66 0.06 0.77 0 0.17 –10 –5 0m1 160 12 –10 –5 0m3 242 12.3 12.7 2.66 0 0.4 0 0 –5 0m3 242 12.6 –10 –5 0m4 242 12.7 2.67 0.13 0.47 0 0.39 –10 –5 0m4 242 12.75 5 10 20m4.1 343.8 12.95 2.68 0.07 0.51 0 0.42 45 80 120m4.1 343.8 13 10 25 40m4.2 364.9 13.1 2.67 0.05 0.56 0 0.34 30 50 80m4.2 364.9 13.12 30 50 80m4.3 377.2 13.2 2.67 0.08 0.24 0 0.67 30 50 80m4.3 377.2 13.22 30 50 80m4.4 409.3 13.28 2.69 0.27 0.07 0 0.66 45 75 110m4.4 409.3 13.3 30 50 80MFS 470.1 13.4 2.69 0.24 0.16 0 0.6 45 75 110m4.5 478.6 13.55 2.65 0.04 0.82 0 0.14 10 25 50m4.5 478.6 13.65 45 75 110m5 502.1 13.75 2.65 0.07 0.56 0 0.37 45 75 110m5 502.1 14.65 40 50 80onlap uncf 511.8 14.8 2.67 0.08 0.51 0 0.33 50 55 80

(continued )

Page 9: Miocene relative sea level on the New Jersey shallow ...geology.rutgers.edu/images/stories/faculty/miller... · onshore and offshore sea-level estimates are offset by epeirogeny,

Research Paper

9Kominz et al. | Miocene relative sea level: IODP 313 New Jersey MarginGEOSPHERE | Volume 12 | Number 5

Stratigraphy above and below the Lower–Middle Miocene

Sediments beneath the Lower to Middle Miocene section sampled by Exp 313 have been compacted under the load of the Miocene and younger sedi-ments. This compaction must be taken into account to properly estimate R1 for the Miocene strata. Based on a contour map of depth to basement (Ben-son, 1984), we estimate that ~8 km of pre-Miocene sediments were deposited beneath the three offshore sites. For modeling purposes, we use the along-strike Anchor Gas, Dickinson No. 1 rotary well (Poag, 1985; Olsson et al., 1988; Sugarman et al., 2011; Fig. 1) to estimate the lithology and age data of the underlying strata.

The Lower to Middle Miocene strata also compacted beneath Upper Mio-cene and younger sediments. These sediments were either poorly sampled (Exp

313 sites M27 and M29) or not sampled at all (site M28). In our model, porosity (and thus compaction) depends only on depth of burial; thus the detailed lithol-ogy and water depth of this younger part of the section have no impact on R1 results. Therefore, the decompaction of the Lower–Middle Miocene focus of the paper is valid because the thickness of overlying strata is known and included in the model. The thermal curve is fit to the R1 water- depth values based on the pre-Miocene Anchor Dickinson sediments overlain by sites M27, M28, or M29 detailed stratigraphic data that are, in turn, overlain by 209, 243, or 342 m of younger strata observed at the latter three sites, respectively. The effects of the younger subsidence history on the magnitudes of Miocene relative sea level are minor, because the R2 results are registered to present sea level, which is unaffected by the intervening upper Miocene and younger section. Corehole water depths at Sites M27, M28, and M29 are 34, 35, and 36 m below present

TABLE 2. AGES, AVERAGE DENSITIES (RHO), AVERAGE LITHOLOGIES, AND WATER DEPTHS* USED AS INPUT FOR SEQUENCES FROM SITES M27, M28, AND M29 (continued )

Depth Age RhoAverage

clayAverage

sandAverageCsand

Averagesilt WD low WD mid WD high

M29 input (continued )

downlap uncf 511.8 14.85 50 80 90MFS 519.6 15.2 2.65 0.08 0.59 0 0.32 75 90 100m5.2 602.2 15.5 2.67 0.18 0.27 0.01 0.54 70 75 100m5.2 602.2 15.75 45 80 120MFS 620.1 15.8 2.66 0.06 0.76 0 0.18 45 80 120m5.3 643.2 16 2.68 0.12 0.48 0 0.37 60 70 80m5.34 643.2 17.6 65 75 100m5.4 662.4 17.7 2.68 0.19 0.19 0 0.57 65 75 100m5.4 662.4 17.75 45 80 120m5.45 683.2 17.85 2.68 0.16 0.34 0 0.47 45 80 120m5.45 683.2 17.9 45 80 120m5.47 687.9 18 2.64 0.14 0.31 0 0.41 45 80 120m5.47 687.9 18.1 45 80 120m5.6 710 18.3 2.67 0.14 0.38 0.01 0.42 45 80 120m5.6 710 18.85 45 80 120m5.7 728.6 19 2.68 0.14 0.4 0 0.42 45 80 120m5.7 728.6 19.45 50 75 100MFS 734.8 19.5 2.68 0.93 0.07 0 0 75 90 110m5.8 746 19.65 2.69 0.88 0.05 0 0 45 75 85m5.8 746 20.6 65 75 100m6 755.5 20.9 2.64 0.16 0.36 0 0.48 65 75 100

Note: Ages and water depths are explicitly defined at the depth indicated and relate to a specific surface (sequence boundary[s], or transgressive [trans.], onlap or maximum flooding surface [MFS]). Only dated surfaces are identified in this table. Sequence boundaries each have two ages, the age above the surface that marks the end of the sequence boundary and the start of deposition of the overlying sequence at this location. Average lithologies given are those of the sequence above this surface and below the overlying dated surface. The age below each sequence boundary marks the end of deposition of the underlying sequence. Detailed lithology data are provided in Supplemental Table S1 (see footnote 1). Shaded rows indicate data that were included in the backstripping, but these data are not discussed because they are not included in the focus of this manuscript. WD—local water depth. Csand—carbonate sand or calcarenite. Abbreviations in first column refer to sequence surfaces shown in Figure 2 and are from Browning et al. (2013) and Mountain et al. (2010). Additionally, uncf—unconformity, Trans Surf—transgressive surface, holo—Holocene strata, PH—undifferentiated Pleistocene and Holocene, P—Pleistocene and/or Pliocene; UlP—upper lower Pliocene, mio—undifferentiated Miocene strata above the detailed section, ba—base of detailed measured corehole, eo—Eocene sediments at base of corehole.

*References cited are Browning et al. (2013), Katz et al. (2013), McCarthy et al. (2013), Ando et al. (2014), Mountain et al. (2010), and Miller et al. (2013a).

Page 10: Miocene relative sea level on the New Jersey shallow ...geology.rutgers.edu/images/stories/faculty/miller... · onshore and offshore sea-level estimates are offset by epeirogeny,

Research Paper

10Kominz et al. | Miocene relative sea level: IODP 313 New Jersey MarginGEOSPHERE | Volume 12 | Number 5

TABLE 3. AGES OF SURFACES FOR SHELF SEQUENCES AND AGES OF STRATA FOR ONSHORE SEQUENCES

Sequence surface M27 age M27 JVB M28 age M28 JVB M29 age M29 JVB Kw age Onshore sequence

m4.1 12.96 13.2 ? 13 ?m4.2 13 ? 13.1 13.1 13.1m4.2 13.1 13.1 13.1 13.12 13.1 13.1 Kw3bm4.3 13.2 13.2 13.2 13.2 13.2 Kw3bm4.3 13.2 13.22 13.2 Kw3bm4.4 ? 13.28 13.3 Kw3bm4.4 13.4 ? 13.36 ? 13.3 13.3 Kw3bMFS 13.4 13.4 Kw3bm4.5 13.5 13.5 13.55 13.3 13.55 13.6 13.5 Kw3bm4.5 13.65 13.6 13.56 13.5 13.65 13.6m5 13.75 13.7 13.7 13.7 13.75 13.7

13.8 Kw3a14.2 Kw3a

m5 14.8 14.8 14.65 14.8 14.65 14.6onlap uncf 14.85 14.85 14.8downlap uncf 14.9 14.9 14.85

15 Kw2bMFS 15.2 15.2 15.2 Kw2bm5.2 15.3 15 15.4 15.1 15.5 15.6 Kw2bm5.2 15.6 15.6 15.7 15.7 15.75 15.8 Kw2bMFS 15.8 15.8 15.8 Kw2b

15.9 Kw2bm5.3 15.95 15.8 16 16.3 16 16

16.1 Kw2am5.3 16.65 16.5 16.5 16.6 Kw2adownlap uncf 16.55 Kw2adownlap uncf 16.6 Kw2aMFS 16.7 16.7 Kw2am5.33 16.75 16.6 16.8 16.7 Kw2am5.33 17.1 16.9 17.05 17.4 Kw2aMFS 17.3 17.3 Kw2am5.34 17.35 17 17.45 17.6 17.5 Kw2am5.34 17.65 17.7 17.6 17.6 17.6 17.6MFS 17.67 17.67m5.4 17.7 17.7 17.7 17.7 17.7m5.4 17.79 ? 17.79 17.9 17.75 17.7m5.45 17.85 18 17.85 18 17.85 17.8m5.45 17.86 17.9 17.9 17.9 17.9 Kw1cm5.47 17.9 18 18 18 Kw1cm5.47 18.1 ? 18.1 18.1 Kw1cm5.6 18.3 18.3 18.3 18.3 Kw1cm5.6 18.6 18.85 18.6 18.85 18.6 Kw1cm5.7 18.8 19 18.8 19 18.8 18.9 Kw1cm5.7 19.1 19.2 19.4 19.5 19.45 20 19.2 Kw1bMFS 19.5 19.5 19.5 Kw1bTrans Surf 19.7 Kw1am5.8 20.1 20.1 19.7 20 19.65 20.2 Kw1a

20.4 Kw1am5.8 20.7 20.7 20.5 20.5 20.6 20.5m6 20.9 20.9 20.9 20.9

Note: Ages are given in Ma (million years before present) for all dated offshore surfaces. These include age assumptions of this work (age) and the ages suggested by Browning et al., 2013 (JVB). All sequence boundaries are given an age for the top of the boundary and the base of the boundary, showing the duration of the unconformity at each site. Maximum flooding surfaces (MFS) are assumed to be correlative where they are observed. Minimum and maximum ages are provided for the composite, onshore Kirkwood (Kw) sequences. The sequence names are indicated when those sequences are present during deposition of offshore surfaces. Abbreviations in the first column refer to the sequence surfaces shown in Figure 2 and are from Browning et al. (2013) and Mountain et al. (2010). Additionally, uncf—unconformity, MFS—maximum flooding surface, and TransSurf—transgressive surface.

Page 11: Miocene relative sea level on the New Jersey shallow ...geology.rutgers.edu/images/stories/faculty/miller... · onshore and offshore sea-level estimates are offset by epeirogeny,

Research Paper

11Kominz et al. | Miocene relative sea level: IODP 313 New Jersey MarginGEOSPHERE | Volume 12 | Number 5

sea level, respectively (Mountain et al., 2010). The registry to present sea level is a reasonable first assumption; although recent GIA effects on the order of 20 m may complicate absolute sea-level estimates (Raymo et al., 2011) and are taken into account in our discussion in the Results section.

Coastal Plain Relative Sea-Level Curve

In this paper, the time scale used by Kominz et al. (2008) to derive a sea-level curve from Miocene sequences has been updated to GTS2012 (Gradstein et al., 2012). The offshore results are of particular interest in comparison to the coastal plain relative sea-level estimates. However, during the two decades these drill core samples have been studied, revisions in the time scale—from Berggren et al. (1995) (BKSA95) to Lourens et al. (2004) and from the GTS2004 to the GTS2012 (Gradstein et al., 2012)—have resulted in small but significant adjustments to the Early to Middle Miocene. Early to Middle Miocene ages are largely unchanged from BKSA95 (used for onshore sequences by previous studies; e.g., Kominz et al., 2008) and offshore by Mountain et al. (2010) from GTS2004 (used by Browning et al., 2013), except for the age of the Oligocene/Miocene boundary. The Early to Middle Miocene time scale of GTS2004 dif-fers from GTS2012 only by placement of the Langhian/Burdigalian boundary (0.16 m.y. older in GTS2012). All coreholes used in generating the coastal plain R2 curve have fairly well defined age ranges for the Miocene sequences, with age errors generally ±0.5 m.y. In sediments younger than 15 Ma, the errors can be larger (i.e., ages are constrained by Sr isotopes that have an error of ±0.75 to ±1.0 m.y.; Browning et al., 2013, their fig. 3). In downdip onshore coreholes, the age ranges are better constrained than they are at updip locations. We used the better dated sequences to constrain the less well dated coreholes. This generally resulted in age shifts of less than 0.5 m.y. and slightly greater amplitudes in the sea-level esti mates, because incidences of destructive inter-ference due to age mis corre lations have been reduced. The Cape May Zoo site (Sugarman et al., 2007) was analyzed subsequent to the Kominz et al. (2008) compilation, and it has been included in our backstripped data set (see Re-sults section). Lithol ogies at all coastal plain sites were decompacted using the coastal plain porosity versus depth relations of Van Sickel et al. (2004).

RESULTS

Coastal Plain Relative Sea-Level Curve

Cape May Zoo R1 and R2

The Cape May Zoo corehole (Sugarman et al., 2007) provides additional Lower to Middle Miocene data that were not included in previous backstrip-ping studies (Kominz et al., 2008). Younger sequences at this site are poorly dated, and they are not included in our R2 analysis. The Oligocene and Eocene portion of the R1 curve (Fig. 4A) is based on sediment from the Cape May Site (Miller and Snyder, 1997), with older units based on the Anchor Dickinson

well (Poag, 1985; Kominz et al., 1998). The beginning of thermal subsidence is taken as 140 Ma, the time when offshore sediment loading overcame lateral thermal uplift landward of the hinge zone (Steckler et al., 1988). For the Middle Miocene section, the subsidence rate is slightly greater than that of the thermal curve, suggesting rising relative sea level during this interval and falling rela-tive sea level between deposition of these Miocene sediments and the present (Fig. 4B).

Combined Results, Coastal Plain Coreholes

Several coreholes drilled over the past decade in the New Jersey coastal plain penetrate Lower and Middle Miocene strata, locally called the Kirkwood Formation. These include coreholes at Ancora (Miller et al., 1999), Bass River (Miller et al., 1998b), and Island Beach (Miller et al., 1994a), where only a few of the Kirkwood sequences were sampled and dated (Kw1a and Kw1b and possi-bly Kw2a or b) and coreholes at Ocean View (Miller et al., 2001), Millville (Sugar-man et al., 2005), Cape May (Miller et al., 1996b), and Atlantic City (Miller et al., 1994b) that sampled multiple Kirkwood sequences. Early and Middle Miocene R2 curves for these sites, in addition to the new Cape May Zoo site (Sugarman et al., 2007), show that several Kirkwood sequences are well represented (Fig. 5A). In previous studies, the timing of sequences did not always match from one corehole to the next due to uncertainties in assigning ages to sequences, particularly in updip locations (e.g., the Ancora and Bass River sites). This re-sulted in a blurring of the sequences so that sequence boundaries were not seen as hiatuses in the combined record (e.g., Kominz et al., 2008). While it is possible that the poor match is real, in most cases the uncertainty of age dates is ~1 m.y. and larger in updip sites. We assume that the coastal plain sequence boundaries are regionally correlated, and we accept the ages of the best dated sequences and adjust the other coreholes accordingly. Seismic correlations on-shore and offshore confirm the correlations (Monteverde et al., 2008; Iscimen, 2014). In general, the Cape May Zoo sequences are well dated and provide ages for the base of most sequences. However, for Kw1 and Kw2a, the sequences at Ocean View provide the best age constraints. Sequences Kw0, Kw1c, and Kw3a are best represented at the Cape May corehole. The tuned sequence ages allow us to stack sequences from several coreholes; the composite distinguishes indi-vidual sequences in the relative sea-level record (Fig. 5B).

The composite relative sea-level curve (Fig. 5C; data are provided in Sup-plemental Table S34, section A) was generated following the procedures of Kominz et al. (2008). The minimum possible range of sea-level estimates is obtained as the minimum high-end sea-level estimate and the maximum low-end sea-level estimate. The best estimate relative sea level is taken as the average of the high- and low-end estimates. Note that despite age model tuning, we have not been able to separate sequences Kw1a and Kw1b that show no discernable hiatus at Ocean View or other coreholes. Nonetheless, we have separated sequences Kw3c into two sequences and separated Kw3a from Kw3b and Kw1b from Kw1c. Error ranges have been reduced in some

TABLE S3B: RELATIVE SEA LEVEL (RSL) AVERAGED NEW JERSEY COASTAL PLAIN. ALL DATA IS IN METERSObserved Observed Observed Shi�ed Shi�ed Shi�ed

Age (Ma) RSL NJCP low RSL NJCP best RSL NJCP high RSL NJCP low RSL NJCP best RSL NJCP high20.420.320.220.120.019.919.819.719.619.519.419.319.2

-33.414-3.395-1.445-5.600-3.170-6.199

-15.018-20.136-11.070-10.358-14.388-18.182-22.676

-6.3589.9118.7764.5883.693

-3.884-8.294

-18.967-7.520-7.719

-10.293-12.477-14.935

20.69923.21718.99714.77710.557-1.569-1.569

-17.797-3.971-5.081-6.199-6.772-7.194

6.58636.43838.22233.90036.16332.96723.98218.69727.59728.14223.94619.98515.324

33.64349.74448.44344.08843.02735.28230.70619.86731.14630.78128.04025.69023.065

60.69963.05158.66454.27749.89037.59837.43121.03634.69533.41932.13531.39530.806

4Supplemental Table S3. Relative sea-level curve for New Jersey Shelf and relative sea-level curve for New Jersey Coastal Plain. Please visit http:// dx .doi .org /10 .1130 /GES01241 .S4 or the full-text article on www .gsapubs .org to view Supplemental Table S3.

Page 12: Miocene relative sea level on the New Jersey shallow ...geology.rutgers.edu/images/stories/faculty/miller... · onshore and offshore sea-level estimates are offset by epeirogeny,

Research Paper

12Kominz et al. | Miocene relative sea level: IODP 313 New Jersey MarginGEOSPHERE | Volume 12 | Number 5

sequences as a result of matching high and low relative sea-level curves, and the timing of the sequences has shifted slightly, mainly due to changes in the geologic time scale.

Exp 313 Offshore Sites M27, M28, and M29

A critical test of the efficacy of one-dimensional backstripping for esti-mating relative sea level is the degree to which the calculated magnitude of R2 is consistent among boreholes. In most instances and within errors of water-depth estimates, our study passes this test for the three Exp 313 drill sites (Fig. 6). Although using low- and high-end porosity versus depth curves increases ranges of uncertainty, the form of m.y.-scale relative sea-level vari-ations remains unchanged (see Supplemental Fig. S25). Sequences younger than sequence boundary m4 (ca. 12.6 Ma) are included in the modeling, but they are not the focus of this study (see Table 2). Poor age and paleoenviron-ment control for these younger sequences could explain the lack of agreement

between sites (Fig. 6). To compare these new offshore results with our relative sea-level curve generated using coastal plain boreholes, we construct a rela-tive sea-level curve based only on the offshore sites, following the method of Kominz et al. (2008) described above. That is, the error ranges are obtained by taking the minimum high estimate and the maximum low estimate of R2. Where multiple coreholes sample the same sequence, this results in a reduc-tion in the range of error bars, because only the overlapping magnitudes are taken to be valid (Fig. 7A; results are provided in Supplemental Table S3 (sec-tion B, [footnote 4]). The “best estimate” R2 curve is based on an average of best estimates for all coreholes that sample sediment of that age. Where R2 results from multiple coreholes do not overlap, the best estimate average does not always fall between the high and low ranges (e.g., m5.3 and m5.2; Fig. 7). In these cases, the error range has been extended to include the best estimate averages (Fig. 7B). Despite these inconsistencies, there is an overall good age agreement, with all three offshore coreholes showing similar R2 trends and magnitudes (Figs. 6 and 7). This suggests that we are reconstructing relative sea-level change at the New Jersey shallow shelf.

0

25

50

75

-25

Rel

ativ

e S

ea L

evel

(m)

Age (m.y.)

m4.2

15.020.0 12.517.5

m4.5

m4.4m4.3m5m5.2m5.3

m5.34m5.7m5.8

Seismic Sequence Boundaries

Seismic Sequences

m5.8 m5.4 m5.2m5.3

m5 m4

m5.6

m5.5m5.6

m4.1m5.4

m5.33

m5.5

M27 M29M28

best estimatehigh porosity

low porosity

Burdigalian SerrvallianLanghian

5Supplemental Figure S2. Relative sea-level (R2) curves from Integrated Ocean Drilling Program Ex-pedition 313 coreholes. Please visit http:// dx .doi .org /10 .1130 /GES01241 .S5 or the full-text article on www .gsapubs .org to view Supplemental Figure S2.

0

300

200

100

Dep

th (m

)

Age (m.y.)20140 040

PPOlig. Miocene

Kw3cKw3b Kw3d

6080120 100

600

500

400

900

800

700

Kw2bKw2aKw1cKw1b

20 18 10121416Age (m.y.)

Rel

ativ

e S

ea L

evel

(m)

R1-sed

50

25

-25

0

Cape May Zoo

Thermal model fits

Cape May Zoo

Paleo.Late Cretaceous EoceneEarly Cretaceous

AB

B

Figure 4. (A) Cape May Zoo backstripping results. R1 curves based on Cape May Zoo: R1 values (Equation 2) older than the Mio-cene are taken from the nearby Anchor Dickinson well (Fig. 1) that was drilled to basement. Unconformities are indicated by gaps in R1 curves. Three R1 curves are generated based on high, low (thin orange lines) and best estimate (thick red line) paleoenvironmental estimates. Also shown is the R1 result when no paleo water depths are taken into account (green curved line labeled “R1-sed”). Each of the four R1 curves is fit to a thermal plate model. These best-fit curves are the smooth, continuous curves of the same color as the R1 curves (see text for discus-sion). (B) Cape May Zoo R2 relative sea-level curves. The difference between R1 and thermal plate model curves corrected for water loading (Equation 3). Only the well-dated Cape May Zoo relative sea level (R2) results are shown. These are part of the Kirkwood (Kw) Formation, as indicated in the figure. PP—Pliocene–Pleistocene; Olig.—Oligocene; Paleo.—Paleocene.

Page 13: Miocene relative sea level on the New Jersey shallow ...geology.rutgers.edu/images/stories/faculty/miller... · onshore and offshore sea-level estimates are offset by epeirogeny,

Research Paper

13Kominz et al. | Miocene relative sea level: IODP 313 New Jersey MarginGEOSPHERE | Volume 12 | Number 5

–25

0

25

Kw1a

Kw1b

Kw2a Kw2b

Kw3a Kw3cKw3b

Kw1c

Atlantic City

Ancora

Ocean View

Cape May Zoo

Bass River

Cape May Millville

Coastal plain

Kw2a

0

25

50

Kw1b

Kw2b

Kw3a Kw3cKw3b

Kw1c

B

–25

C

Kw1a

Kw3a - Kw3c

–25

0

25

50

A

Kw1a & Kw1b

Kw2a Kw2bKw1c

15.0

Burdigalian SerrvallianLanghian

Age (m.y.)20.0 12.517.5

Rel

ativ

e S

ea L

evel

(m)

Rel

ativ

e S

ea L

evel

(m)

Rel

ativ

e S

ea L

evel

(m)

Figure 5. (A) Relative sea-level (R2) curves from Miocene sequences in coastal plain coreholes. Each corehole is indicated by its color in the key between 5A and 5B. Mio-cene stages are indicated on the top. The Early/Middle Miocene boundary occurs between the Burdigalian and Langhian Stages. (B) New R2 values for Miocene onshore sequences. Ages of Miocene se-quences were adjusted so that the best dated sequences constrain the ages of less well-dated correlative sequences. (C) Composite Miocene relative sea-level curve from New Jersey coastal plain core-holes using data plotted in panel 5B (see Supplemental Table S3 (section B) for coastal plain sea-level data [footnote 4]).

Page 14: Miocene relative sea level on the New Jersey shallow ...geology.rutgers.edu/images/stories/faculty/miller... · onshore and offshore sea-level estimates are offset by epeirogeny,

Research Paper

14Kominz et al. | Miocene relative sea level: IODP 313 New Jersey MarginGEOSPHERE | Volume 12 | Number 5

DISCUSSION

The relative sea-level curve generated from offshore data shows higher magnitudes of sea-level fluctuations than the onshore data (Fig. 8A). Taking into account the full error range, the largest long-term sea-level range allowable in the shelf data is 75 m (16.7–13 Ma) compared to a 55 m range in the onshore data (20.4–20.3 Ma). On the other hand, the minimum allowable sea-level ranges of the two data sets are 32 m in the offshore data (16.7–12.2 Ma) and a 17 m in the onshore data (19.2–13.4 Ma). These ranges are entirely compatible with those observed in comparable time intervals at the Marion Plateau by John et al. (2011). They report four sequences with maximum sea-level changes of 65 m and minimum ranges of 26 m between 13.8 and 16.6 Ma. All estimates are lower than the 100–120 m variations implied by the global compilation of Haq et al. (1987), but are compatible with magnitudes of 20–50 m suggested by Haq and Al-Qahtani (2005) for the Arabian Peninsula.

Comparing the averaged R2 curves from Exp 313 with the revised compos-ite onshore relative sea-level curve has implications for models of tectonics and GIA effects (Fig. 8A). From a sequence stratigraphic approach, we would ex-pect coastal plain sequence deposition to represent the most proximal portion of transgressive and/or highstand systems tracts. These might occur during slowly rising relative sea level that may be associated with non-deposition off-shore. Thus, coastal plain sequences would be present during the maximum amplitude of relative sea level. Our initial comparison of onshore and offshore sites (onshore is shown as light gray units in Fig. 8A) shows that deposition on the coastal plain sometimes occurs during hiatuses in the corresponding offshore sedimentation sampled by Exp 313. Where this is the case, it implies that the sequence boundaries of offshore units indicate relative sea-level rise coupled with offshore sediment starvation rather than the more commonly ac-cepted model of sea-level fall (e.g., Loutit et al., 1988). In most cases, onshore and offshore sequences show significant temporal overlap, with the offshore

m4.2m4.5

m4.3m5.7m5.8 m5.6 m5.4 m4.4

m5m5.2m5.3m5.33

M27

M29

M28

0

25

50

75

100

–25

Rel

ativ

e S

ea L

evel

(m)

Age (m.y.)15.020.0 12.517.5

Seismic Sequence Boundaries

Seismic Sequences

Burdigalian SerrvallianLanghian

m5.8

m5.4

m5.2

m5.3m5

m4m5.5

m5.6

m5.7

Figure 6. Relative sea-level (R2) curves from Integrated Ocean Drilling Program (IODP) Expedition 313 coreholes. Time missing across sequence boundaries is shown by yellow rectangles along the bot-tom. The range of low, high, and best esti-mate R2 values of depositional sequences (color coded by well in key at the upper right) vary from very close to modern sea level (0 m) to roughly 80 m above modern. The sequences are labeled above the R2 values. Miocene stages are indicated at the top of the graph. The Early/Middle Miocene boundary occurs between the Burdigalian and Langhian Stages.

Page 15: Miocene relative sea level on the New Jersey shallow ...geology.rutgers.edu/images/stories/faculty/miller... · onshore and offshore sea-level estimates are offset by epeirogeny,

Research Paper

15Kominz et al. | Miocene relative sea level: IODP 313 New Jersey MarginGEOSPHERE | Volume 12 | Number 5

0

25

50

75

Rel

ativ

e S

ea L

evel

(m)

Age (m.y.)

m4.2

15.020.0 12.517.5

m4.5

m4.4m4.3

m5m5.2m5.3m5.34

m5.7m5.8

Seismic Sequence Boundaries

m5.6m4.1

m5.4m5.33

Burdigalian SerrvallianLanghianBurdigalian SerrvallianLanghian

M27

M29

M28

Seismic Sequencesm5.8

m5.4

m5.2

m5.3m5

m4

m5.5

m5.6

m5.7

Lowest highHighest lowAverage bestRange overlap

0

25

50

–25

Average R2with error

ranges

B

A

Figure 7. (A) Average relative sea level (bold purple line) superimposed on the R2 estimates shown in Figure 6. This average is based on all best estimate R2 values from the lowest maximum estimate of sea level to the highest minimum estimate of sea level at any time among the three wells. (B) Average relative sea level (bold purple line) derived from Exp 313 corehole data similar to 7A, but this time based on the average of all best estimates and the error range derived from results in Figure 7A (see text; see Supplemental Table S3 (section A) for tabulated data [see footnote 4]). Sequence boundaries, sequences, and the Miocene time scale are labeled as in Figure 6.

Page 16: Miocene relative sea level on the New Jersey shallow ...geology.rutgers.edu/images/stories/faculty/miller... · onshore and offshore sea-level estimates are offset by epeirogeny,

Research Paper

16Kominz et al. | Miocene relative sea level: IODP 313 New Jersey MarginGEOSPHERE | Volume 12 | Number 5

0

25

50

75

Rel

ativ

e S

ea L

evel

(m)

Age (m.y.)15.020.0 12.517.5

Seismic Sequence Boundariesm4.2m4.5

m4.3m5.7m5.8 m5.6 m5.4 m4.4

m5m5.2m5.3m5.33

Burdigalian SerrvallianLanghianBurdigalian SerrvallianLanghian

Seismic Sequencesm5.8

m5.4m5.2m5.3

m5m5.5

m5.6m5.7

0

25

50

–25

Average R2with error

ranges

B

A

adjustedcoastal plain

coastal plain

Average R2with error

ranges

e ssm

Kw3cKw3bKw3a

Kw2bKw2a

Kw1a

Kw1c

Kw1b

Kw3c

Kw3b

Kw3aKw2b

Kw2aKw1a

Kw1cKw1b

Figure 8. (A) Relative sea-level (R2) curves for both the coastal plain composite relative sea-level curve (from Fig. 5C) and R2 average results from the Inte-grated Ocean Drilling Program (IODP) Expedition 313 coreholes (from Fig. 7B). (B) Best estimate coastal plain composite sea-level curve corrected for non- thermal motion of the coastal plain relative to the shelf is plotted in dark gray for comparison with best estimate offshore sea-level curve. See text for discus-sion. Sequence boundaries, sequences, and the Miocene time scale are labeled as in Figure 6.

Page 17: Miocene relative sea level on the New Jersey shallow ...geology.rutgers.edu/images/stories/faculty/miller... · onshore and offshore sea-level estimates are offset by epeirogeny,

Research Paper

17Kominz et al. | Miocene relative sea level: IODP 313 New Jersey MarginGEOSPHERE | Volume 12 | Number 5

generally more complete. There are similar trends in offshore and onshore data in overlapping intervals. In fact, sequences Kw3b and Kw3c are, at least in part, compatible with relative sea level obtained from analyzing offshore strata (Fig. 8A). Thus, in most cases, both onshore and offshore sequence bound-aries represent sea-level falls.

The R2 sea-level estimates obtained from the onshore sites (0 ± 20 m above present sea level) are considerably lower than those indicated by the offshore sites (40 ± 30 m above present sea level), particularly in the older Miocene units. This is in direct contrast with stratigraphic principles that require sea-level rise to be higher when proximal land is flooded (e.g., Posamentier and Vail, 1988). As such, it suggests that the New Jersey margin has experienced non-thermal epeirogeny, resulting in differential subsidence of the onshore and offshore portions of the margin. The dominant epeirogenic effect predicted to have oc-curred in this region is the overriding of the Farallon plate. As the cold, dense lithosphere of the subducted slab passes beneath the margin, coastal New Jersey experiences subsidence (Conrad et al., 2004). Moucha et al. (2008) pre-dicted that over the past 30 m.y., the coastal plain of New Jersey has subsided 50–100 m more than offshore New Jersey. The minimum value of 50 m of dif-ferential subsidence over 30 m.y. implies a relative coastal plain uplift of 33 m at 20 Ma, reducing to 20 m at 12 Ma. We add this differential subsidence to the onshore relative sea-level curve to make it comparable to the offshore R2 results. This correction for epeirogenic effects largely reconciles differences between onshore and offshore relative sea-level estimates.

An additional variation between onshore and offshore relative sea level is predicted by the effect of ice loading (GIA). Raymo et al. (2011) estimated the impact of ice volume and GIA effects on observations of Pliocene sea-level maxima. They found that GIA effects have diminished through time for all but the last glacial maximum and the very near-field location of the melted ice sheets. However, the impact of the last glacial maximum on the elevations of the New Jersey coastal plain and the New Jersey shelf estimated by Raymo et al. (2011) will also affect older sea-level estimates by distorting today’s datum. They found that the New Jersey coastal plain yields relative sea-level estimates that are ~12 m above eustatic sea level, while the shelf yields relative sea-level estimates that are ~18–24 m above eustatic sea level (depending on distance offshore and on the assumptions made for lithosphere rheology). Therefore, offshore relative sea-level estimates will always appear to be 6–12 m higher than coastal plain estimates. Thus, to compare onshore relative sea-level maxima with offshore data, 6–12 m needs to be added to all onshore data. Rowley et al. (2013) also suggest five to ten meters of relative uplift occurred between the coastal plain and shelf as a result of GIA, con-sistent with the Raymo et al. (2011) results. The effect of the Farallon slab in conjunction with the GIA effects shift relative sea-level variations predicted by the coastal plain curve relative to the offshore R2 curve (Fig. 8B). The resulting coastal plain adjusted-R2 magnitudes, in comparison with the offshore data, are consistent with the assumption that the deposition on the onshore coastal plain generally occurs at or near the highest magnitude of sea level. Our cor-

rected relative sea-level curves for onshore and offshore New Jersey provide a working model for eustatic changes for the Early to Middle Miocene.

It is premature to draw broad conclusions about the implications of these results for the relationship between sequences and systems tracts for se-quence paradigms because of limitations in the backstripping technique and because of uncertainties in correlating sequences from onshore to offshore. Such detailed comparisons are hampered by the fact that onshore data are compiled from seven widely dispersed locations, while the offshore data set consists of three coreholes along a single dip transect.

Correlation of the data sets relies on both chronostratigraphy and the basic assumptions of sequence stratigraphy (including superposition). All strata above a sequence boundary are younger than the strata below; this is also true for maximum flooding surfaces. This allows correlation of sequences amongst onshore locations and offshore locations with finer resolution than provided by the chronostratigraphy (e.g., we are reasonably certain that sequence Kw1a is correlatable throughout the coastal plain coreholes). We are also reasonably certain of the onshore to offshore correlation not only through chronostratigra-phy but also through pattern matching and seismic correlations. For example, the onshore composite sequence Kw2a sequence can be precisely correlated with composite sequence m5.4 using well logs and seismic profiles (Iscimen, 2014). Thus, though the uncertainty on the numerical age of any given back-stripped estimate is large (±0.25 to ±0.5 m.y.), physical correlations (principles of superposition, seismic control, and sequence stratigraphy) mean that our relative age correlations amongst sites is finer.

If deposition did not follow the tenets of sequence stratigraphy, then cor-relation would rely solely on chronostratigraphy. The age uncertainty on the ages of sequences is large (±0.25–0.5 m.y. offshore and ±0.5–1.0 m.y. onshore) compared to the mean sea-level cycle duration of 0.84 ± 0.52 m.y. Exp 313 also sampled several higher-order sequences (100/405 k.y.; Browning et al., 2013); thus the average duration of our sequences is shorter than 1 m.y. Never the-less, our sequences are of a higher order than predicted by changes in mantle dynamic topography that appear to operate on longer-term time scales (2–100 m.y.; Petersen et al., 2010). Reliance on chronostratigraphy alone would not alter the pervasively lower relative magnitude of sea level recorded on the coastal plain as compared to that observed on the shelf. That is, relative subsidence of the coastal plain would still be required, supporting the epeiro-genic model. We expect more precise relative sea-level interpretations when the onshore and offshore core data are combined with seismic imaging in a full two-dimensional backstripping approach (e.g., Kominz and Pekar, 2001). Future work will include additional correlations of offshore sequences using seismic profiles and well logs (e.g., Iscimen, 2014) and two-dimensional back-stripping (e.g., Steckler et al., 1999).

In summary, comparison of R2 results from the new offshore coreholes with the revised coastal plain sequences requires relative subsidence of the coastal plain since the Middle Miocene. This is compatible with modeling of mantle-driven subsidence due to the subducted Farallon slab (e.g., Moucha et al., 2008).

Page 18: Miocene relative sea level on the New Jersey shallow ...geology.rutgers.edu/images/stories/faculty/miller... · onshore and offshore sea-level estimates are offset by epeirogeny,

Research Paper

18Kominz et al. | Miocene relative sea level: IODP 313 New Jersey MarginGEOSPHERE | Volume 12 | Number 5

CONCLUSIONS

Backstripped relative sea-level estimates of Middle to Late Miocene se-quences from the New Jersey coastal plain and three new IODP shelf sites are generally internally consistent with respect to the timing of sequence bound-aries and relative sea-level variations. Coastal plain sedimentation tends to predict relative sea-level changes that are consistent with those seen offshore. Where onshore and offshore sedimentation corresponds, the rising and falling portions of the relative sea-level curve can be correlated, although the magni-tude of offshore and onshore estimates is offset. This result requires that the New Jersey passive margin has undergone epeirogeny and, in particular, that the offshore shelf has subsided less than the coastal plain. Thus, it is consistent with relative epeirogeny due to subduction of the Farallon plate (Moucha et al., 2008) and due to GIA effects of the last deglaciation (Raymo et al., 2011). The amplitude of Late to Middle Miocene m.y.-scale sea-level changes seen at the New Jersey margin is generally 5–20 m but occasionally is as great as 50 m.

ACKNOWLEDGMENTS

This manuscript was greatly improved by suggestions from one anonymous reviewer, reviewers Hugo Pouderoux and Olivier Dauteuil, and Geosphere Associate Editor Jean-Noël Proust. This work was supported by funds from the Faculty Research and Creative Activities Award, Western Michigan University to Kominz. Support was provided from National Science Foundation grants EAR-1052257, OCE-1154379, and OCE14-63759 to Miller. Funding was supplied by the U.S. Sci-ence Support Program Consortium for Ocean Leadership to Katz, Miller, Browning, and Mountain. Samples were provided by the IODP.

REFERENCES CITED

Ando, H., Oyama, M., and Nanayama, F., 2014, Data report: Grain size distribution of Miocene successions, IODP Expedition 313 Sites M0027, M0028, and M0029, New Jersey shallow shelf, in Mountain, G., Proust, J.-N., McInroy, D., Cotterill, C., and the Expedition 313 Scien-tists, Proceedings of the Integrated Ocean Drilling Program 313: Tokyo, Integrated Ocean Drilling Program Management International, Inc., doi: 10 .2204 /iodp .proc .313 .201 .2014 .

Benson, R.N., 1984, MS2 Structure contour map of Pre-Mesozoic basement, landward margin of Baltimore Canyon Trough (North and South Sections): Delaware Geological Survey Miscel-laneous Map 2, scale 1:500,000.

Berggren, W.A., Kent, D.V., Swisher, C.C., III, and Aubry, M.-P., 1995, A revised Cenozoic geochro-nology and chronostratigraphy, in Berggren, W.A., Kent, D.V., Aubry, M.-P., and Harde, J., eds., Geochronology, Time Scales, and Global Stratigraphic Correlations: SEPM (Society for Sedimentary Geology) Special Publication 54, p. 129–212, doi: 10 .2110 /pec .95 .04 .0129 .

Bond, G.C., and Kominz, M.A., 1984, Construction of tectonic subsidence curves for the early Paleozoic miogeocline, southern Canadian Rocky Mountains: Implications for subsidence mechanisms, age of breakup, and crustal thinning: Geological Society of America Bulletin, v. 95, p. 155–173, doi: 10 .1130 /0016 -7606 (1984)95 <155: COTSCF>2 .0 .CO;2 .

Bond, G.C., Kominz, M.A., and Grotzinger, J.P., 1988, Cambro-Ordovician eustasy: Evidence from geophysical modeling of subsidence in Cordilleran and Appalachian passive margins, in Paola, C., and Kleinspehn, K., eds., New Perspectives in Basin Analysis: New York, Springer- Verlag, p. 129–160, doi: 10 .1007 /978 -1 -4612 -3788 -4_7 .

Bond, G.C., Kominz, M.A., Steckler, M.S., and Grotzinger, J.P., 1989, Role of thermal subsidence, flexure and eustasy in the evolution of early Paleozoic passive-margin carbonate platforms, in Crevello, P.D.,Wilson, J.L., Sarg J.F., and Read, J.F., eds., Controls on Carbonate Platform and Basin Development, SEPM (Society for Sedimentary Geology) Special Publication, v. 44, p. 39–61, doi: 10 .2110 /pec .89 .44 .0039 .

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–Miocene sequences on the New Jersey shallow shelf: Implications for regional, interregional, and global correlations: Geo-sphere, v. 9 , p. 1434–1456, doi: 10 .1130 /GES00857 .1 .

Conrad, C., Lithgow-Bertelloni, C., and Louden, K.E., 2004, Iceland, the Farallon slab, and dy-namic topography of the North Atlantic: Geology, v. 32, p. 177–180, doi: 10 .1130 /G20137 .1 .

Grabau, A.W., 1936, Oscillation or pulsation: 16th International Geological Congress Report, v. 1, p. 539–553.

Gradstein, F.M., Ogg, J.G., Schmidtz, M.D., and Ogg, G.M., eds., 2012, The geologic time scale: New York, Elsevier, 1144 p.

Haq, B.U., and Al-Qahtani, A.M., 2005, Phanerozoic cycles of sea-level changes on the Arabian Platform: Geoarabia, v. 10, p. 127–160.

Haq, B.U., Hardenbol, J., and Vail, P.R., 1987, Chronology of fluctuating sea levels since the Tri-assic (250 million years ago to Present): Science, v. 235, p. 1156–1167, doi: 10 .1126 /science .235 .4793 .1156 .

Haq, B.U., Hardenbol, J., and Vail, P.R., 1988, Mesozoic and Cenozoic chronostratigraphy and cycles in sea level change, in Wilgus, C.K., Hastings, B.S., Kendall, C.G.St.C., Posamentier, H.W., Ross, C.A., and Van Wagoner, J.C., eds., Sea-Level Changes: An Integrated Approach: SEPM (Society for Sedimentary Geology) Special Publication 42, p. 71–108, doi: 10 .2110 /pec .88 .01 .0071 .

Iscimen, T., 2014, Sequence stratigraphy of Miocene sequences Kw2a and m5.4, New Jersey: Onshore to offshore correlations [M.S. thesis]: New Brunswick, New Jersey, Rutgers Uni-versity, 80 p.

John, C.M., Karner, G.D., Browning, E., Leckie, R.M., Mateo, Z., Carson, B., and Lowery, C., 2011, Timing and magnitude of Miocene eustasy derived from the mixed siliciclastic-carbonate stratigraphic record of the northeastern Australian margin: Earth and Planetary Science Let-ters, v. 304, p. 455–467, doi: 10 .1016 /j .epsl .2011 .02 .013 .

Katz, M.E., Miller, K.G., and Mountain, G.S., 2003, Biofacies and lithofacies evidence for paleo-environmental interpretations of upper Neogene sequences on the New Jersey continental shelf (ODP Leg 174A), in Olson, H.C., and Leckie, R.M., eds., Micropaleontologic Proxies for Sea-Level Change and Stratigraphic Discontinuities: SEPM (Society for Sedimentary Geol-ogy) Special Publication 75, p. 131–146, doi: 10 .2110 /pec .03 .75 .0131.

Katz, M.E., Browning, J.V., Miller, K.G., Monteverde, D., Mountain, G.S., and Williams, R.H., 2013, Paleobathymetry and sequence stratigraphic interpretations from benthic foraminifera: In-sights on New Jersey shelf architecture, IODP Expedition 313: Geosphere, v. 9, p. 1488–1513, doi: 10 .1130 /GES00872 .1 .

Kominz, M.A., and Pekar, S.F., 2001, Oligocene eustasy from two-dimensional sequence strati-graphic backstripping: Geological Society of America Bulletin, v. 113, p. 291–304, doi: 10 .1130 /0016 -7606 (2001)113 <0291: OEFTDS>2 .0 .CO;2 .

Kominz, M.A., Miller, K.G., and Browning, J.V., 1998, Long-term and short-term global Ceno-zoic sea-level estimates: Geology, v. 26, p. 311–314, doi: 10 .1130 /0091 -7613 (1998)026 <0311: LTASTG>2 .3 .CO;2 .

Kominz, M.A., Browning, J.V., Miller, K.G., Sugarman, P.J., Misintzeva, S., and Scotese, C.R., 2008, Late Cretaceous to Miocene sea-level estimates from the New Jersey and Delaware coastal plain coreholes: An error analysis: Basin Research, v. 20, p. 211–226, doi: 10 .1111 /j .1365 -2117 .2008 .00354 .x .

Kominz, M.A., Patterson, K., and Odette, D., 2011, Lithology dependence of porosity in slope and deep marine sediments: Journal of Sedimentary Research, v. 81, p. 730–742, doi: 10 .2110 /jsr .2011 .60 .

Kopp, R.E., Mitrovica, J.X., Griffies, S.M., Yin, J., Hay, C.C., and Stouffer, R.J., 2010, The impact of Greenland melt on local sea levels: A partially coupled analysis of dynamic and static equi-librium effects in idealized water-hosing experiments: Climatic Change, v. 103, p. 619–625, doi: 10 .1007 /s10584 -010 -9935 -1 .

Lourens, L., Hilgen, F., Shackleton, N.J., Laskar, J., and Wilson, D., 2004, The Neogene Period, in Gradstein, F., Ogg, J., and Smith, A., eds., A Geologic Time Scale: Cambridge University Press, p. 409–440.

Page 19: Miocene relative sea level on the New Jersey shallow ...geology.rutgers.edu/images/stories/faculty/miller... · onshore and offshore sea-level estimates are offset by epeirogeny,

Research Paper

19Kominz et al. | Miocene relative sea level: IODP 313 New Jersey MarginGEOSPHERE | Volume 12 | Number 5

Loutit, T.S., Hardenbol, J., Vail, P.R., and Baum, G.R., 1988, Condensed section: The key to age determination and correlation of continental margin sequences, in Wilgus, C.K., Hastings, B.S., Kendall, C.G.St.C., Posamentier, H.W., Ross, C.A., Van Wagoner J.C., eds., Sea-Level Changes: An Integrated Approach: SEPM (Society for Sedimentary Geology) Special Publi-cation 42, p. 183–213, doi: 10 .2110 /pec .88 .01 .0183 .

McCarthy, F.M.G., Katz, M.E., Kotthoff, U., Drljepan, M., Zanatta, R., Williams, R.H., Browning, J.V., Hesselbo, S.P., Bjerrum, C., Miller, K.G., and Mountain, G.S., 2013, Eustatic control of New Jersey margin architecture: Palynological evidence from IODP Expedition 313: Geo-sphere, v. 9, p. 1457–1487, doi: 10 .1130 /GES00853 .1 .

McKenzie, D., 1978, Some remarks on the development of sedimentary basins: Earth and Plane-tary Science Letters, v. 40, p. 25–32, doi: 10 .1016 /0012 -821X (78)90071 -7 .

Miller, K.G., and Snyder, S.W., eds., 1997, Proceedings of the Ocean Drilling Program, Scientific Results, Leg 150X: College Station, Texas, Ocean Drilling Program, 151 p.

Miller, K.G., Wright, J.D., and Fairbanks, R.G., 1991, Unlocking the ice house: Oligocene-Mio-cene oxygen isotopes, eustasy and margin erosion: Journal of Geophysical Research, v. 96, p. 6829–6848, doi: 10 .1029 /90JB02015 .

Miller, K.G., Browning, J.V., Liu, C., Sugarman, P., Kent, D.V., and Van Fossen, M., Queen, D., Goss, M., Gwynn, D., Mullikin, L., Feigenson, M.D., Aubry, M.-P., and Burckle, L.D., 1994a, Atlantic City site report, in Miller, K.G., et al., ed., Proceedings of the Ocean Drilling Program, Initial reports, Volume 150X: College Station, Texas, Ocean Drilling Program, p. 35–55.

Miller, K.G., Sugarman, P., Van Fossen, M., Liu, C., Browning, J.V., Queen, D., Aubry, M.-P., Burckle, L.D., Goss, M., and Bukry, D., 1994b, Island Beach site report, in Miller, K.G., et al., eds., Proceedings of the Ocean Drilling Program, Initial Reports, Volume 150X: College Sta-tion, Texas, Ocean Drilling Program, p. 5–33.

Miller, K.G., Mountain, G.S., and Leg 150 Shipboard Party Members of the New Jersey Coastal Plain Drilling Project, 1996a, Drilling and dating New Jersey Oligocene–Miocene sequences: Ice volume, global sea level, and Exxon records: Science, v. 271, no. 5252, p. 1092–1095, doi: 10 .1126 /science .271 .5252 .1092 .

Miller, K.G., Liu, C., Browning, J.V., Pekar, S.F., Sugarman, P.J., Van Fossen, M.C., Mullikin, L., Queen, D., Feigenson, M.D., Aubry, M.-P., Burckle, L.D., Powars, D., and Heibel, T., 1996b, Cape May site report, in Miller, K.G., et al., ed., Proceedings of the Ocean Drilling Program, Initial reports, Volume 150X (Supplement): College Station, Texas, Ocean Drilling Program, p. 1–28.

Miller, K.G., Browning, J.V., Pekar, S.F., and Sugarman, P.J., 1997, Cenozoic evolution of the New Jersey coastal plain: Changes in sea level, tectonics, and sediment supply, in Miller, K.G., and Snyder, S.W., eds., Proceedings of the Ocean Drilling Program, Scientific results, Vol-ume 150X: College Station, Texas, Ocean Drilling Program, p. 361–373, doi: 10 .2973 /odp .proc . sr .150X .326 .1997 .

Miller, K.G., Mountain, G.S., Browning, J.V., Kominz, M., Sugarman, P.J., Christie-Blick, N., Katz, M.E., and Wright, J.D., 1998a, 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., Sugarman, P.J., Browning, J.V., Olsson, R.K., Pekar, S.F., Reilly, T.R., Cramer, B.S., and Aubry, M.-P., Lawrence, R.P., Curran, J., Stewart, M., Metzger, J.M., Uptegrove, J., Bukry, D., Burckle, L.H., Wright, J.D., Feigenson, M.D., Brenner, G.J., and Dalton, R.F., 1998b, Bass River Site Report, Proceedings of the ODP, Initial Reports, 174AX: College Station, Texas, Ocean Drilling Program, 39 p.

Miller, K.G., Sugarman, P.J., Browning, J.V., Cramer, B.S., Olsson, R.K., de Romero, L., Aubry, M.-P., Pekar, S.F., Georgescu, M.D., Metzger, K.T., Monteverde, D.H., Skinner, E.S., Uptegrove, J., Mullikin, L.G., Muller, F.L., Feigenson, M.D., Reilly, T.J., Brenner, G.J., and Queen, D., 1999, Ancora Site, in Miller, K.G., Sugarman, P.J., Browning, J.V., et al., eds., Proceedings of the Ocean Drilling Program, Initial Reports, Volume 174AX (Supplement): College Station, Texas, Ocean Drilling Program, p. 1–65.

Miller, K.G., Sugarman, P.J., Browning, J.V., Pekar, S.F., Katz, M.E., Cramer, B.S., Monteverde, D., Uptegrove, J., McLaughlin, P.P., Jr., Baxter, S.J., Aubry, M.-P., Olsson, R.K., Van Sickel, B., Metzger, K., Feigenson, M.D., Tiffin, S., and McCarthy, F., 2001, Ocean View Site, in Miller, K.G., Sugarman, P.J., Browning, J.V., et al., eds., Proceedings of the Ocean Drilling Program, Initial Reports, Volume 174AX (Supplement): College Station, Texas, Ocean Drilling Pro-gram, p. 1–72.

Miller, K.G., Kominz, M.A., Browning, J.V., Wright, J.D., Mountain, G.S., Katz, M.E., Sugarman, P.J., Cramer, B.S., Christie-Blick, N., and Pekar, S.F., 2005, The Phanerozoic record of sea-level change: Science, v. 310, p. 1293–1298, doi: 10 .1126 /science .1116412 .

Miller, K.G., Mountain, G.S., Wright, J.D., and Browning, J.V., 2011, A 180-million-year record of sea level and ice volume variations from continental margin and deep-sea isotopic records: Washington, D.C., Oceanography, v. 24, p. 40–53, doi: 10 .5670 /oceanog .2011 .26 .

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., 2013a, Sequence boundaries are impedance contrasts: Core-seismic-log integration of Oligocene-Miocene sequences, New Jersey shallow shelf: Geo-sphere, v. 9, p. 1257–1285, doi: 10 .1130 /GES00858 .1 .

Miller, K.G., Mountain, G.S., Browning, J.V., Katz, M.E., Monteverde, D., Sugarman, P.J., Ando, H., Bassetti, M.A., Bjerrum, C.J., Hodgson, D., Hesselbo, S., Karakaya, S., Proust, J.-N., and Rabineau, M., 2013b, Testing sequence stratigraphic models by drilling Miocene foresets on the New Jersey shallow shelf: Geosphere, v. 9, p. 1236–1256, doi: 10 .1130 /GES00884 .1 .

Milne, G.A., Gehrels, W.R., Hughes, C.W., and Tamisiea, M.E., 2009, Identifying the causes of sea-level change: Nature Geoscience, v. 2, p. 471–478, doi: 10 .1038 /ngeo544 .

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 .

Moucha, R., Forte, A.M., Mitrovica, J.X., Rowley, D.B., Quéré, S., Simmons, N.A., and Grand, S.P., 2008, Dynamic topography and long-term sea-level variations: There is no such thing as a stable continental platform: Earth and Planetary Science Letters, v. 271, p. 101–108, doi: 10 .1016 /j .epsl .2008 .03 .056 .

Mountain, G., Proust, J.-N., McInroy, D., and Cotterill, C., and the Expedition 313 Scientists, 2010, Proceedings of the Integrated Ocean Drilling Program, 313: Tokyo, Integrated Ocean Drilling Program Management International, Inc., 515 p., doi: 10 .2204 /iodp .proc .313 .101 .2010 .

Müller, D., Sdrolias, M., Gaina, C., Steinberger, B., and Heine, C., 2008, Long-term sea level fluc-tuations driven by ocean basin dynamics: Science, v. 319, p. 1357–1362, doi: 10 .1126 /science .1151540 .

Olsson, R.K., Gibson, T.G., Hansen, H.J., and Owens, J.P., 1988, Geology of the northern Atlantic coastal plain: Long Island to Virginia, in Sheridan, R.E., and Grow, J.A., eds., The Atlantic Continental Margin, U.S.: Boulder, Colorado, Geological Society of America, Geology of North America, v. I-2, p. 87–105.

Parker, F.L., 1948, Foraminifera of the continental shelf from the Gulf of Maine to Maryland: Har-vard Museum of Comparative Zoology Bulletin, v. 100, p. 213–241.

Peltier, W.R., 1998, Postglacial variations in the level of the sea: implications for climate dynamics and solid-Earth geophysics: Reviews of Geophysics, v. 36, p. 603–689, doi: 10 .1029 /98RG02638 .

Petersen, K.D., Nielsen, S.B., Clausen, O.R., Stephenson, R., and Gerya, T., 2010, Small-scale mantle convection produces stratigraphic sequences in sedimentary basins: Science, v. 329, p. 827–830, doi: 10 .1126 /science .1190115 .

Poag, C.W., 1981, Ecologic atlas of benthic foraminifera of the Gulf of Mexico: Stroudsburg, Pennsylvania, Hutchison Ross Publishing Co., 174 p.

Poag, C.W., 1985, Depositional history and Stratigraphic reference section for central Baltimore Canyon Trough, in Poag, C.W., ed., Geologic Evolution of the United States Atlantic Margin: New York, Van Nostrand Reinhold, p. 217–263.

Posamentier, H.W., and Vail, P.R., 1988, Eustatic controls on clastic deposition II—Sequence and sys-tems tract models, in Wilgus, C.K., Hastings, B.S., Kendall, C.G.St.C., Posamentier, H.W., Ross, C.A., and Van Wagoner, J.C., eds., Sea-Level Changes: An Integrated Approach: SEPM (Society for Sedimentary Geology) Special Publication 42, p. 125–154, doi: 10 .2110 /pec .88 .01 .0125 .

Posamentier, H.W., Jervey, M.Y., and Vail, P.R., 1988, Eustatic controls on clastic deposition I— Sequence and systems tract models, in Wilgus, C.K., Hastings, B.S., Kendall, C.G.St.C., Posa-mentier, H.W., Ross, C.A., and Van Wagoner, J.C., eds., Sea-Level Changes: An Integrated Approach: SEPM (Society for Sedimentary Geology) Special Publication 42, p. 109–124, doi: 10 .2110 /pec .88 .01 .0109 .

Raymo, M.E., Mitrovica, J.X., O’Leary, M.J., DeConto, R.M., and Hearty, P.J., 2011, Departures from eustasy in Pliocene sea-level records: Nature Geoscience, v. 4, p. 328–332, doi: 10 .1038 /ngeo1118 .

Rowley, D.B., Forte, A.M., Moucha, R., Mitrovica, J.X., Simmons, N.A., and Grand, S.P., 2013, Dynamic topography change of the eastern United States since 3 million years ago: Science, v. 340, p. 1560–1563, doi: 10 .1126 /science .1229180 .

Royden, L., and Keen, C.E., 1980, Rifting process and thermal evolution of the continental margin of Eastern Canada determined from subsidence curves: Earth and Planetary Science Letters, v. 51, p. 343–361, doi: 10 .1016 /0012 -821X (80)90216 -2 .

Page 20: Miocene relative sea level on the New Jersey shallow ...geology.rutgers.edu/images/stories/faculty/miller... · onshore and offshore sea-level estimates are offset by epeirogeny,

Research Paper

20Kominz et al. | Miocene relative sea level: IODP 313 New Jersey MarginGEOSPHERE | Volume 12 | Number 5

Spasojević, S., Liu, L., Gurnis, M., and Müller, R.D., 2008, The case for dynamic subsidence of the United States East Coast since the Eocene: Geophysical Research Letters, v. 35, L08305, doi: 10 .1029 /2008GL033511 .

Steckler, M.S., and Watts, A.B., 1978, Subsidence of the Atlantic-type continental margin off New York: Earth and Planetary Science Letters, v. 41, p. 1–13, doi: 10 .1016 /0012 -821X (78)90036 -5 .

Steckler, M.S., Watts, A.B., and Thorne, J.A., 1988, Subsidence and basin modeling at the U.S. Atlantic passive margin, in Sheridan, R.E., and Grow, J.A., eds., The Atlantic Continental Margin: Boulder, Colorado, Geological Society of America, Geology of North America, v. I-2, p. 399–416.

Steckler, M.S., Mountain, G.S., Miller, K.G., and Christie-Blick, N., 1999, Reconstruction of Ter-tiary progradation and clinoform development on the New Jersey passive margin by 2-D backstripping: Marine Geology, v. 154, p. 399–420, doi: 10 .1016 /S0025 -3227 (98)00126 -1 .

Sugarman, P.J., Miller, K.G., Browning, J.V., McLaughlin, P.P., Jr., Brenner, G.J., Buttari, B., Cramer, B.S., Harris, A., Hernandez, J., Katz, M.E., Lettini, B., Misintseva, S., Monteverde, D.H., Olsson, R.K., Patrick, L., Roman, E., Wojtko, M.J., Aubry, M.-P., Feigenson, M.D., Barron, J.A., Curtin, S., Cobbs, G., Cobbs, G., III, Bukry, D., and Huffman, B., 2005, Millville Site, in Miller, K.G., Sugar-

man, P.J., Browning, J.V., et al., eds., Proceedings of the Ocean Drilling Program, Initial reports, Volume 174AX (Supplement): College Station, Texas, Ocean Drilling Program, p. 1–94.

Sugarman, P.J., Miller, K.G., Browning, J.V., Monteverde, D.H., Uptegrove, J., McLaughlin, J.P.P., Stanley, A.M., Wehmiller, J., Kulpecz, A., Harris, A., Pusz, A., and Kahn, A., 2007, Cape May Zoo Site, in Miller, K.G., Sugarman, P.J., Browning, J.V., et al., eds., Proceedings of the Ocean Drilling Program, Initial reports, Volume 174AX (Supplement): College Station, Texas, p. 1–66.

Sugarman, P.J., Monteverde, D.H., Pristas, R., Girard, M., Boyle, J., Miller, K.G., Browning, J.V., Fan Reinfelder, Y., Romero, P., and Kulpecz, A., 2011, Characterization of the carbon dioxide storage potential beneath the New Jersey Coastal Plain: Preliminary Characterization of CO2 Sequestration Potential in New Jersey and the Offshore Coastal Region, p. 1–44, http:// www .mrcsp .org /userdata /phase _ii _reports /njgs _carbon _sequestration _report _web .pdf.

van Morkhoven, F.P.C.M., Berggren, W.A., Edwards, A.S., 1986, Cenozoic cosmopolitan deep- water benthic foraminifera: Pau, France, Elf-Aquitaine, Memoire 11, 421 p.

Van Sickel, W.A., Kominz, M.A., Miller, K.G., and Browning, J.V., 2004, Late Cretaceous and Ceno zoic sea-level estimates: Backstripping analysis of borehole data, onshore New Jersey: Basin Research, v. 16, p. 451–465, doi: 10 .1111 /j .1365 -2117 .2004 .00242 .x .