<<

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

GEOSPHERE Miocene relative sea level on the New Jersey shallow continental shelf and coastal plain derived from one-dimensional GEOSPHERE; v. 12, no. 5 backstripping: A case for both eustasy and epeirogeny doi:10.1130/GES01241.1 M.A. Kominz1, K.G. Miller2, J.V. Browning3, M.E. Katz4, and G.S. Mountain2 8 figures; 3 tables; 5 supplemental files 1Department of Geosciences, Western Michigan University, 1186 Rood Hall, 1903 West Michigan Avenue, Kalamazoo, Michigan 49008, USA 2Department 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 CORRESPONDENCE: michelle.kominz@​ wmich​ .edu​ 3Department of Earth and Planetary Sciences, Rutgers, The State University of New Jersey, 610 Taylor Road, Piscataway, New Jersey 08854-8066, USA 4Department of Earth and Environmental Sciences, Rensselaer Polytechnic Institute, 110 8th Street, Troy, New York 12180, USA 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-­ ABSTRACT et al., 1988). We use the term relative sea-level change to include changes in dimensional backstripping: A case for both eustasy­ ­eustasy coupled with epeirogenic (broad regional uplift; Grabau, 1936) and and epeirogeny: Geosphere, v. 12, no. 5, p. 1– 20, doi:10.1130/GES01241.1. Onshore drilling by Ocean Drilling Program (ODP) Legs 150X and 174AX ­local changes in the height of lithosphere relative to the center of the Earth. and offshore drilling by Integrated Ocean Drilling Program (IODP) Expedition That is, relative sea-level change is measured relative to a fixed point on the Received 31 July 2015 313 provides continuous cores and logs of seismically imaged Lower to ­Middle crust (Posamentier et al., 1988). Eustasy is of particular importance because Revision received 9 May 2016 Miocene sequences. We input ages and paleodepths of these sequences into it serves as the datum against which Earth’s tectonic and climate history can Accepted 3 August 2016 one-dimensional backstripping equations, progressively accounting for the be measured (e.g., Miller et al., 2005). In this paper, we attempt to untangle effects of compaction, Airy loading, and thermal subsidence. The resulting ­thermal subsidence and epeirogeny from eustasy. difference between observed subsidence and theoretical thermal subsidence The timing and magnitude of relative sea level has been studied in detail provide relative sea-level curves that reflect both global average sea level and at the Late to Miocene of the New Jersey margin. Onshore New non-thermal subsidence. In contrast with expectations, backstripping sug- Jersey drilling by Ocean Drilling Program (ODP) Legs 150X and 174AX in con- gests that the relative sea-level maxima in proximal onshore sites were lower cert with Legs 150 and 174A outer shelf and continental slope drilling (Fig. 1; than correlative maxima on the shelf. This requires that the onshore New Jer- e.g., Miller et al., 1997, 1998a; Mountain et al., 2010) has shown that the timing sey coastal plain has subsided relative to the shelf, which is consistent with of sequence boundaries (erosional surfaces recognized in cores and seismic models of relative epeirogeny due to of the . These profiles), is consistent withd 18O increases from deep-sea records (Miller et al., models predict subsidence of the coastal plain relative to the shelf. Although 1991, 1996a, 2005, 2011; Browning et al., 2008). This suggests that relative sea- onshore and offshore sea-level estimates are offset by epeirogeny, the ampli­ level changes observed on the New Jersey margin were caused, at least in tude of million-–scale Early to sea-level changes seen at part, by eustatic variations due to ice growth and decay. the New Jersey margin is generally 5–20 m and occasionally as great as 50 m. Backstripping is a modeling technique that accounts for the effects of sedi­ These events are interpreted to represent eustatic variations, because they oc- ment compaction, sediment loading, and, in this case, thermal subsidence. cur on a shorter time frame than epeirogenic influences. Correction for epeiro­ Applied to the onshore coreholes, backstripping provides a relative sea-level genic effects largely reconciles differences between onshore and offshore rela­ record for the New Jersey margin that in the absence of other tectonic effects tive sea-level estimates and suggests that backstripping provides a testable yields a testable estimate of eustatic change (Miller et al., 2005; Kominz et al., eustatic model for the Early to Middle Miocene. 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 (e.g., Conrad et al., 2004; INTRODUCTION Moucha et al., 2008; Spasojević et al., 2008). Thus, more work is required to separate eustatic and epeirogenic effects in this region. One of the outstanding challenges in studying Earth history is document- Several processes must be accounted for to untangle eustasy from epeirog­ ing the timing and magnitude of global sea-level change (e.g., Haq et al., 1987, eny. Most backstripping estimates of the magnitude of sea-level change in this For permission to copy, contact Copyright 1988; Miller et al., 2005). Here we use eustasy to mean the global change in region have used a one-dimensional approach that assumes an Airy response Permissions, GSA, or [email protected]. sea level relative to a fixed point, e.g., the center of the Earth (Posamentier to sediment loads (Kominz et al., 1998; Miller et al., 1998a, 2005; Van Sickel

© 2016 Geological Society of America

GEOSPHERE | Volume 12 | Number 5 Kominz et al. | Miocene relative sea level: IODP 313 New Jersey Margin 1 Research Paper

–75° –74° –73° –72°

P L F O H E R S C A L T N T U N E O T I N C S O I U C O O Z E C E O A N Sea Figure 1. Location map. Sites used in this E T E L E C Girt work are indicated as large red circles for A R O P C I 40° M M27 Integrated Ocean Drilling Program (IODP) M28 Expedition 313 sites and as large green cir- Island M29 Beach cles for the onshore coreholes with lower Ancora to middle Miocene strata. Seismic profiles Ft. Mott Bass River 1071 are from three different data acquisition New cruises (R/V Ewing cruise Ew9009, R/V Jersey 1072 1073 Oceanus cruise Oc270, and R/V Cape City ­Hatteras cruise CH0698; Monteverde et al., Millville 2008; Mountain et al., 2010; Miller et al., Ocean View 2013a). The seismic section Oc270 line Cape 529 (red line passing through Exp 313 drill May sites) is shown in Figure 2. The Anchor Zoo Anchor 906 Dickinson well (gray filled dot; Poag, 1985; 00 39° Dickinson * 20 well Seismic Profiles Sugarman et al., 2011) was drilled as a gas a Cape 902 exploration well between the Cape May CH0698 Oc270 903 May 0 904 and Cape May Zoo sites. AMCOR—Atlan- ula 10 0 Ew9009 Figure 2 0 10 tic Margin Coring Project; DSDP—Deep Sea Drilling Program. DrillDrillsites Sites 905 Bethany Beach IODP Expedition 313 offshore ODP onshore ODP 200 DSDP AMCOR 00 N oil exploration 30

et al., 2004; Kominz et al., 2008). Only one model, of Upper to lower- ing of glaciers (glacial isostatic adjustment [GIA]) has been shown to vary most Miocene strata, used a two-dimensional backstripping approach that in- globally (e.g., Peltier, 1998) with impact in this region (e.g., Raymo et al., 2011). corporates flexural rigidity of the lithosphere to account for subsidence caused This effect is most pronounced during the large Northern Hemisphere ice ages by sediment loads at a distance (Kominz and Pekar, 2001). There are several of the past 2.7 m.y., but GIA influences the reference frame of older records as complications inherent in estimates of New Jersey margin relative sea level. well (e.g., Raymo et al., 2011). One issue is the fact that coastal plain sediments rarely contain a complete Drilling data provided by IODP Expedition 313 (hereafter “Exp 313”) on record of sea-level change. They generally preserve only the transgressive the New Jersey shallow continental shelf focused on Lower to Middle Mio- and highstand systems tracts, leaving lowstand sediments farther seaward cene strata (Mountain et al., 2010). This data set presents an opportunity to beneath what is now the continental shelf (Miller et al., 1998a). As discussed estimate amplitudes of offshore relative sea-level change based on cores, above, other tectonic effects have been postulated in this region so that tec- logs, and seismic profiles that can be compared to the onshore results. By tonic subsidence may not be entirely thermal. In particular, the arrival of the providing estimates of the magnitude of relative sea-level change, we can Farallon slab beneath the North American east coast ca. 75 Ma requires that begin to address the magnitude and timing of both eustasy and more regional New Jersey subsided beyond the predicted thermal subsidence (Conrad et al., epeirogeny. 2004; Kominz et al., 2008; Moucha et al., 2008; Müller et al., 2008; Spasojević Exp 313 drilled three coreholes (sites M27, M28, and M29; Fig. 1) in ~30 m et al., 2008). This means that the stratigraphic succession in this region has of water, targeting Miocene sequences that also were cored in multiple loca- been imprinted by eustatic, thermal, and epeirogenic processes. Additionally, tions on the onshore coastal plain. The Miocene section is well imaged on a the whole Earth response to loading effects of water coupled with the unload- grid of seismic profiles that show a of clinothems, which are packages

GEOSPHERE | Volume 12 | Number 5 Kominz et al. | Miocene relative sea level: IODP 313 New Jersey Margin 2 Research Paper

of sediment that prograde seaward and are bounded by surfaces (in this case, METHODS AND INPUT DATA sequence boundaries) with distinct sigmoidal (clinoform) geometries (Fig. 2; Mountain et al., 2010). These sites, drilled as a transect along seismic line Backstripping utilizes compaction, age, and paleodepth observations from Oc270 529, provide a two-dimensional cross section of several stratigraphic cores or outcrops to estimate how the basement would have subsided (tec- sequences formed between 12 Ma and 22 Ma (Mountain et al., 2010). Here, tonic subsidence [TS]) in the absence of sediments and eustatic sea-level we present the results of a one-dimensional backstripping study of this new change (DSL) (Steckler and Watts, 1978; Bond and Kominz, 1984): data set. As such, it is directly comparable to the one-dimensional modeling  ρm −ρs  ρw  already published from the coastal plain (Kominz et al., 2008). By extending TS =ΦS*   −∆SL   −∆SL +WD, (1) ρm −ρw ρm −ρw 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 where: r is density of the asthenospheric mantle (m), the decompacted sedi­ consider the effects of the Farallon slab and GIA on relative sea-level change ment (s), and seawater (w); WD is local water depth; S* is the decompacted across the New Jersey margin. sediment thickness; and F is the basement response function.

IODP Expedition 313 27 28 29 0 NW SE 0

.1 .1

m4.1 .2 m5 .2

m5.2 .3 m1 .3

m5.45 m5.3 m3 Figure 2. Seismic profile Oc270 529. Seis- s m5.32 mic interpretations from Monteverde et al. .4 m5.6 .4 m5.4 m4 (2008) and Mountain et al. (2010). Deposi- tional sequences are highlighted in vari­ ous colors and named according to the

Second underlying sequence boundary. For exam- .5 m5.8 .5 m6 m5.33 ple, at M27 the strata colored blue rest on m5.7 m4.4 sequence boundary m5.4 and belong to m5.34 m4.3 m5.47 sequence m5.4. IODP—Integrated Ocean .6 o.5 m4.2 .6 K/T m4.1 Drilling Program. o1 m5 m4.5 .7 m5.6 .7 m5.2

.8 m5.3 .8

m5.4 .9 .9

1.0 1.0 10000 9000 8000 7000 6000 5000 4000 3000 2000 1000 cdp Oc270 529 010 km

GEOSPHERE | Volume 12 | Number 5 Kominz et al. | Miocene relative sea level: IODP 313 New Jersey Margin 3 Research Paper

Table S-1. Porosity vs. Depth Curves In practice, backstripping begins with measuring porosity as a function of ery (Bond et al., 1988). Therefore, the difference between R1 and thermal

Clay best Porosity = 82.7e-depth/418 <172 m depth and lithology in a borehole or outcrop. These data are used to estimate tectonics, R2, is not affected by one-dimensional backstripping (Kominz Clay best Porosity = 61.4 e-depth/1676 >172 m Clay low Porosity = 45 e-depth/417 < 172 m the porosity, and thus, decompacted thickness (S*) of every part of the sedi- et al., 1998). What is lost through one-dimensional analysis of a flexurally Clay low Porosity = 33 e-depth/1674 > 172 m mentary column through time. The sediment column is removed at each time subsiding thermal basin is the detailed relationship between data points Clay high Porosity = 90 e-depth/425 < 95 Clay high Porosity = 76 e-depth/1674 > 95 step and replaced by a column of seawater that represents the hole that sedi­ along a two-dimensional profile that could add insights for interpretations

Silt best Porosity = 75 e-depth/419 < 97m ment of the estimated porosity, thickness, and density would have filled. In (e.g., Steckler et al., 1999; Kominz and Pekar, 2001). Observed relative sea Silt best Porosity = 63 e-depth/1675 > 97m Silt low Porosity = 60 e-depth/420 < 80 calculating the effect of the sediment load, we have applied a one-dimensional level is also complicated by glacial isostatic adjustment (GIA; Peltier, 1998), Silt low Porosity = 52 e-depth/1676 > 80 approach that assumes the underlying lithosphere has no rigidity (the Airy by gravitational, rotational, and flexural effects due to changing ice sheets Silt high Porosity = 85 e-depth/418 < 63 Silt high Porosity = 76 e-depth/1677 > 63 model, in which F, in Equation 1 is taken as 1). We also account for changes (collectively known as “static equilibrium” effects; Kopp, et al., 2010) and by

Sand best Porosity = 54.5 e-depth/1648 in water depth as described below. For that, paleodepth estimates (WD) are (e.g., Milne et al., 2009). These issues will be consid- -depth/2164 Sand low Porosity = 40 e added to the water column that was filled with sediment (above) to determine ered in evaluating the results. Sand high Porosity = 70 e-depth/3566 the total subsidence of the basement in the absence of sediment. The result is Carbonate best Porosity = 88 e-depth/1338 Carbonate low Porosity = 25 e-depth/3126 termed the first reduction, or R1 (Bond et al., 1989): Carbonate high Porosity = 90 e-depth/1730 Lithology and Porosity  ρm −ρs R1= S *+ WD, (2) ρm −ρw 1Supplemental Table S1. Porosity versus depth curves. We use the backstripping approach of Bond et al. (1989) in which poros­ Please visit http://dx​ .doi​ .org​ /10​ ​.1130/GES01241​ ​.S1 or the full-text article on www​.gsapubs.org​ to view Sup- where the variables are the same as described in Equation 1. R1 does not take ity is assumed to be lithology dependent and the porosity of mixed lithol­ plemental Table S1. into account global average sea-level change ( = eustasy; DSL in Equation 1). ogies is calculated based on the proportion of lithologies present. This R1 is based on observed data that can be obtained from a corehole with a has been shown to be a viable approach for marine sediments (Kominz reasonable degree of accuracy. et al., 2011). Porosities based on moisture and density measurements from Porosity (%) 20 30 40 50 60 70 80 0610 20 30 40 50 0 70 80 90 0 Tectonism at a passive margin may be assumed to follow the theoretical the three sites used in this study were combined with smear slide inter-

100 behavior of a cooling plate due to lithospheric stretching during rifting, as pretations, grain-size analysis of discrete samples, and downhole log data 200 demonstrated by the fact that long-term (10–100 m.y.) passive margin records (Mountain, et al., 2010; Miller et al., 2013a; Ando et al., 2014) to obtain lithol-

300

) can be largely explained with an exponential fit (e.g., McKenzie, 1978; Royden ogy-dependent porosity versus depth plots for the New Jersey shelf (Fig. 3 400 and Keen, 1980). Because of the predictable of tectonics in this setting, and Table 1). Lithologies were extended from the depths of each discrete Depth (m 500 Porosity Data Sandstone we can estimate eustasy by separating the long-term, thermal component of measurement to the nearest stratigraphic boundary (Fig. 2) above and be- Silt 600 Clay subsidence from any perturbations observed in the R1 curve. A second param- low that measurement; otherwise, the lithologic boundaries were placed Porosity Relations Sandstone 700 Silt eter, R2, is calculated as the difference between R1 and a cooling plate model between samples. Siliciclastic sand-, silt- and clay-dominated intervals are Clay 800 0610 20 30 40 50 0 70 80 fit to the observed R1 data. We fit a thermal plate model (following Kominz sufficiently distinct in the Exp 313 cores to provide valid comparisons with et al., 2008) to the R1 curve to determine the thermal component of subsid- standard, lithology-based porosity versus depth plots published elsewhere. 2Supplemental Figure S1. Integrated Ocean Drilling Program Expedition 313 porosity versus depth for ence. The difference between the estimated thermal subsidence (taken as TS, While the Exp 313 data show no consistent pattern with depth, their trends sediments dominated by (>50%) sands, silts, and Equation 1) and R1 (Equation 2), after being corrected for the change in water are close to the >90% clay, the >90% silt, and the >60% sand curves based on clay. Please visit http://dx​ .doi​ .org​ /10​ ​.1130/GES01241​ ​ load (again assuming a one-dimensional Airy response of the underlying litho­ a global Ocean Drilling Program (ODP) database (Kominz et al., 2011). Thus, .S2 or the full-text article on www​.gsapubs.org​ to view Supplemental Figure S1. sphere) yields the second reduction R2: 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  ρw  R2 = (R1−TS). (3) versus depth curves that encompass the range of observed porosities were TABLE S2C: INPUT DATA M29 ρm −ρw Thickness age DensityClayMicrite Sand CaCO 3 SiltWD lowWD midWD high   PH 3.535 0.010 2.705 0.235 0.000 0.0600.000 0.70536.00036.00036.000 1 P3.590 0.070 2.705 0.235 0.000 0.0600.000 0.70510.00025.00040.000 also applied for comparison (see Supplemental Table S1 and Supplemental Pleist 21.755 2.691 0.185 0.000 0.2600.000 0.55510.00025.00040.000 Pleist 32.240 2.661 0.098 0.000 0.6100.000 0.29310.00025.00040.000 2 Pleist 40.545 2.707 0.243 0.000 0.0300.000 0.72810.00025.00040.000 Employing Equations 1 and 2 to define TS and R1 explicitly defines R2 Fig. S1 ). Pleist 50.820 2.710 0.250 0.000 0.0000.000 0.75010.00025.00040.000 Pleist 60.915 2.703 0.235 0.000 0.0600.000 0.70510.00025.00040.000 Pleist 70.620 2.657 0.108 0.000 0.5200.050 0.32310.00025.00040.000 as a eustatic estimate. However, this is only true if all tectonics at the bore- Pleist 81.480 2.632 0.095 0.000 0.5700.050 0.28510.00025.00040.000 P1 0.000 0.090 2.700 0.000 0.000 0.0000.000 0.00010.00025.00040.000 P1 0.135 0.100 2.632 0.095 0.000 0.5700.050 0.2855.000 10.000 20.000 Pleist 9 13.005 2.6110.008 0.000 0.9600.010 0.0235.000 10.000 20.000 hole in question are governed by thermal subsidence. Additional errors Pleist 10 14.360 2.613 0.013 0.000 0.9500.000 0.0385.000 10.000 20.000 P2 0.000 0.1102.700 0.000 0.000 0.0000.000 0.0005.000 10.000 20.000 P2 2.105 1.030 2.613 0.013 0.000 0.9500.000 0.0385.000 10.000 20.000 may arise as a result of assuming Airy isostasy, in a basin that actually Ages and Environments of Deposition Pleist 11 4.895 2.645 0.038 0.000 0.8500.000 0.1135.000 10.000 20.000 UlP0.000 1.080 2.700 0.000 0.000 0.0000.000 0.0005.000 10.000 20.000 responded flexurally to the sediment load because subsidence is under­ 3Supplemental Table S2. Ages, densities (rho), lithol- esti­mated at the locus of loading and overestimated at the periphery of the The focus of this work is the Lower–Middle Miocene section, for which ogies, and water depths used as input for sequences basin. However, in our approach we are looking at the difference between considerable data are available and has been synthesized to provide input from coreholes M27 (section A), M28 (section B), and R1 and the theoretical thermal subsidence fit to the observed subsidence, for modeling (Table 2; detailed input data are provided in Supplemental ­Table M29 (section C). Please visit http://dx​ .doi​ .org​ /10​ ​.1130​ 3 /GES01241​.S3 or the full-text article on www​.gsapubs​ R1. The form of subsidence resulting from the flexural response to loading S2 ). Age estimates are based on Browning et al. (2013), who used calcare- .org to view Supplemental Table S2. of a thermal basin is thermal in both the center of the basin and its periph- ous nannofossil, dinoflagellate cyst, and biostratigraphy, and numer-

GEOSPHERE | Volume 12 | Number 5 Kominz et al. | Miocene relative sea level: IODP 313 New Jersey Margin 4 Research Paper

Porosity (%) TABLE 1. POROSITY VERSUS DEPTH CURVES 0610 20 30 40 50 0 70 80 90 ClayPorosity = 82.7 e–depth/418 <172 m 0 ClayPorosity = 61.4 e–depth/1676 >172 m Porosity Data Silt Porosity = 75 e–depth/419 <97 m 100 Sandstone Silt Porosity = 63 e–depth/1675 >97 m Silt Sand Porosity = 54.5 e–depth/1648 Clay CarbonatePorosity = 88 e–depth/1338 200 Porosity Relations Note: The porosity versus depth curves used in this model are provided above. Sandstone Porosity values, e.g., porosity at the surface and calculated porosity, are in volume 300 Silt percent. Depths, decay constant, and depth ranges are given in meters (m). Porosity Clay 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). 400 Depth (m) 500 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

600 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- 700 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- 800 cific date within the sequence. Finally, where onshore and offshore sequences Figure 3. Porosity versus depth measured in Integrated Ocean Drilling Pro- are correlative, minor shifts in ages (0–0.45 m.y.) of the offshore strata were gram (IODP) Expedition 313 sediments composed of >50% sand (yellow made to align correlative sequences. All assigned dates are within the error circles), >50% silt (green diamonds), or >50% clay-sized particles (brown limits presented by Browning et al. (2013). squares.) The yellow, green, and brown lines show porosity-depth rela- tions from a database of measurements in other ODP boreholes (Kominz Environments of deposition at the three well sites were based on benthic et al., 2011) used for decompacting each sediment type in this study. See foraminiferal assemblages, planktonic foraminiferal abundances, and palyno­ text for discussion and Table 1 for curves used in modeling. Porosity ver- logical proxies (Katz et al., 2013; McCarthy et al., 2013; Miller et al., 2013b) 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 coupled with lithofacies interpretations (Browning et al., 2013). Generally the S1 and Supplemental Figure S1 (footnotes 1 and 2, respectively). 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 ous strontium (Sr) isotopic age estimates of calcium carbonate from mollusk as inner neritic (0–30 m), middle neritic (30–100 m), and outer neritic (100– shells, shell fragments, and foraminiferal tests. Ages were assigned using 200 m) (van Morkhoven et al., 1986). Miller et al. (1997) and Katz et al. (2013) the time scale of Gradstein et al. (2012; GTS2012). By using Sr-isotope stra- used a subdivision of these zones on the New Jersey margin when they in- tigraphy, Browning et al. (2013) overcame the challenges of poor magneto­ terpreted Elphidium-dominated biofacies as <10 m, Hanzawaia hughesi-domi­ stratigraphy and poor biostratigraphy posed by coarse, clastic, nearshore nated biofacies as 10–25 m, Pseudononion pizarrensis–dominated biofacies as sediments. They were able to generate age resolution of typically ±0.5 m.y., 25–50 m, Bulimina gracilis–dominated biofacies as 50–80 m, and Uvigerina-­ and in many sequences, age resolution was as good as ±0.25 m.y., where dominated biofacies as >75 m. For deeper biofacies not recovered by Miller age resolution refers to uncertainties in correlation to the et al. (1997), Katz et al. (2013) used key taxa (e.g., Cibicidoides pachyderma, (Browning et al., 2013). Cibicidoides primulus, Hanzawaia mantaensis, and Oridorsalis umbonatus) Superposition and sequence stratigraphic principles provide additional often found in high-diversity, low-dominance assemblages that indicate outer relative age constraints. Our biostratigraphic and Sr-isotopic data have been neritic paleodepths (100–200 m) (e.g., Parker, 1948; Poag 1981; van Morkhoven combined with seismic profiles, downhole logs, sedimentological data, and et al., 1986; Katz et al., 2003). Because changes from the Uvigerina spp. bio- sequence stratigraphic interpretations to provide age estimates for sequence facies to the deeper biofacies occur gradually within sequences, Katz et al. boundaries and several prominent subsequence stratal surfaces (Browning (2013) inferred that the maximum water depths are in the shallower part of the et al., 2013). In this study, dates of some sequence boundaries and maximum outer neritic zone (~120 m). Water-depth ranges as well as best estimate water flooding surfaces (MFS) have been shifted slightly from previous work of depths are used in modeling.

GEOSPHERE | Volume 12 | Number 5 Kominz et al. | Miocene relative sea level: IODP 313 New Jersey Margin 5 Research Paper

TABLE 2. AGES, AVERAGE DENSITIES (RHO), AVERAGE LITHOLOGIES, AND WATER DEPTHS* USED AS INPUT FOR SEQUENCES FROM SITES M27, M28, AND M29 Average Average Average Average Depth Age Rho clay sand Csand silt WD lowWD mid WD high M27 input holo 0 0.01 2.66 0.07 0.79 0.02 0.1343434 P 0.28 0.011 2.66 0.07 0.79 0.02 0.1343434 P 0.28 0.072 0510 P 13.57 0.085 2.66 0.13 0.65 0.02 0.19 0510 P 13.57 0.19 20 35 55 P 22.8 0.22 2.68 0.26 0.33 0.01 0.39 20 35 55 P 22.8 1.03 0510 P 27.41 1.08 2.66 0.14 0.62 0.02 0.21 0510 P 27.41 1.4 0510 MP 32.93 1.5 2.66 0.13 0.65 0.01 0.20510 MP 32.93 11.4 –5 05 m1 97.03 11.8 2.66 0.13 0.67 00.19–505 m1 97.03 12 –5 05 m3 112.07 12.3 2.67 0.34 0.15 00.51–505 m3 112.07 12.6 –5 05 m4 136.03 12.7 2.68 0.21 0.48 00.31–505 m4.1 136.03 12.96 –5 05 m4.2 180.71 13 2.67 0.18 0.55 00.27510 20 m4.2 184.09 13.1 15 30 50 m4.3 209 13.2 2.69 0.28 0.23 00.48203555 m4.4 209 13.4 15 30 50 m4.5 218.39 13.5 2.69 0.10.330.010.55510 20 m4.5 218.39 13.65 20 35 55 m5 225 13.75 2.66 0.10.560.010.33203555 m5 225 14.8 10 20 35 onlap uncf 227.51 14.85 2.69 0.11 0.34 0.06 0.47 35.1 65.9 97.6 downlap uncf 227.51 14.9 35.1 65.9 97.6 MFS 228 15.2 2.7 0.09 0.14 0.01 0.76 40 75 110 m5.2 236.15 15.3 2.67 0.06 0.46 00.47510 20 m5.2 236.15 15.6 20 35 55 MFS 249.76 15.8 2.7 0.09 0.08 0.02 0.81 25 75 80 m5.3 256.31 15.95 2.7 0.09 0.15 0.02 0.75 15 30 50 m5.3 256.31 16.65 35 45 55 MFS 264.61 16.7 2.7 0.09 0.12 0.01 0.78 35 45 55 m5.33 271.35 16.75 2.69 0.07 0.34 0.01 0.58 25 30 50 m5.33 271.35 17.1 10 15 25 MFS 288.64 17.3 2.69 0.09 0.38 0.01 0.52 35 40 50 m5.34 295.12 17.35 2.69 0.17 0.30.020.52354050 m5.34 295.12 17.65 10 25 40 MFS 297.7 17.67 2.7 0.14 0.10 0.76 05080 m5.4 314.67 17.7 2.7 0.11 0.15 0.01 0.73 10 25 40 m5.4 315.78 17.79 20 40 70 m5.45 336.17 17.85 2.69 0.09 0.29 0.02 0.6254060 m5.45 336.17 17.86 51020 m5.47 355.64 17.9 2.64 0.05 0.83 00.12510 20 (continued )

GEOSPHERE | Volume 12 | Number 5 Kominz et al. | Miocene relative sea level: IODP 313 New Jersey Margin 6 Research Paper

TABLE 2. AGES, AVERAGE DENSITIES (RHO), AVERAGE LITHOLOGIES, AND WATER DEPTHS* USED AS INPUT FOR SEQUENCES FROM SITES M27, M28, AND M29 (continued ) Average Average Average Average Depth Age Rho clay sand Csand silt WD lowWD mid WD high M27 input (continued ) m5.6 355.64 18.6 0515 m5.7 361.39 18.8 2.67 0.13 0.57 00.3 51020 m5.7 361.39 19.1 0510 MFS 434.96 19.5 2.67 0.40.590 04075110 TransSurf 477.79 19.7 2.69 0.86 0.13 00 20 35 55 m5.8 494.98 20.1 2.67 0.28 0.71 00 30 50 80 m5.8 494.98 20.7 40 75 110 m6 515.11 20.9 2.68 0.13 0.53 0.02 0.31 40 75 110 m6 515.1123 40 75 110 o6 538.79 23.5 2.67 0.13 0.55 0.02 0.29 40 75 110 o6 538.79 28.2 40 75 110 o3 617.11 29.3 2.68 0.17 0.40.020.414075110 o3 617.11 32.2 40 75 110 o1 625.94 32.3 2.7 0.21 0.30 0.49 40 75 110 o1 625.94 33.6 30 50 80 ba 629.44 33.8 2.71 0.26 0.03 0.01 0.61 30 50 80 eo 629.44 55 50 70 100 M28 input holo 0 0.01 2.66 0.02 0.90 0.08 35 35 35 Mio 223.7 10 2.7 0.20.010 0.79 25 40 60 m4.2 223.7 13.1 30 50 80 m4.3 244 13.2 2.7 0.20.020 0.79 30 50 80 m4.4 244 13.36 0510 MFS 250.6 13.4 2.65 0.09 0.53 0.01 0.36 20 35 55 m4.5 254.2 13.55 2.7 0.18 0.08 00.73102540 m4.5 254.2 13.56 0510 m5 276.8 13.7 2.66 0.03 0.84 0.02 0.12 10 25 40 m5 276.8 14.65 51025 onlap uncf 303.6 14.85 2.63 0.08 0.56 0.01 0.35 19.3 36.6 54 downlap uncf 303.6 14.9 19.3 36.6 54 MFS 310.2 15.2 2.71 0.10.090 0.81 30 50 70 m5.2 323.2 15.4 2.69 0.12 0.31 00.56102530 m5.2 323.2 15.7 10 15 25 MFS 330.4 15.8 2.55 0.02 0.87 0.01 0.1102540 m5.3 361 16 2.64 0.02 0.91 0.01 0.07 10 25 40 m5.3 361 16.5 51020 downlap uncf 386.2 16.55 2.66 0.02 0.91 00.079.1 22.4 36.6 downlap uncf 386.2 16.6 9.122.436.6 MFS 391 16.7 2.66 0.03 0.85 00.11102540 m5.33 404 16.8 2.65 0.02 0.90.010.06510 20 m5.33 404 17.05 10 25 40 MFS 449 17.3 2.65 0.06 0.66 00.27305080 m5.34 479 17.45 2.65 0.07 0.46 00.47305080 m5.34 479 17.6 40 75 110 (continued )

GEOSPHERE | Volume 12 | Number 5 Kominz et al. | Miocene relative sea level: IODP 313 New Jersey Margin 7 Research Paper

TABLE 2. AGES, AVERAGE DENSITIES (RHO), AVERAGE LITHOLOGIES, AND WATER DEPTHS* USED AS INPUT FOR SEQUENCES FROM SITES M27, M28, AND M29 (continued ) Average Average Average Average Depth Age Rho clay sand Csand silt WD lowWD mid WD high M28 input (continued ) MFS 504.9 17.67 2.68 0.08 0.35 00.574075110 m5.4 512.3 17.7 2.65 0.03 0.71 00.26305080 m5.4 512.3 17.79 40 75 110 m5.45 533.6 17.85 2.65 0.07 0.58 00.354075110 m5.45 533.6 17.9 40 75 110 m5.47 545.5 18 2.65 0.05 0.83 00.124075110 m5.47 545.5 18.1 40 75 110 m5.6 567.5 18.3 2.66 0.04 0.87 00.094075110 m5.6 567.5 18.85 40 75 110 m5.7 611.6 19 2.66 0.06 0.81 00.134075110 m5.7 611.6 19.4 10 25 40 MFS 654.3 19.5 2.71 0.87 0.13 00 45 50 75 m5.8 663 19.7 2.66 0.53 0.47 00 45 50 75 m5.8 663 20.5 30 50 80 m6 668.6 20.9 2.65 0.23 0.23 00.54305080 M29 input PH 0 0.01 2.71 0.24 0.06 00.71363636 P 3.54 0.07 10 25 40 P1 15.5 0.09 2.68 0.17 0.29 0.01 0.52 10 25 40 P1 15.5 0.1 51020 P2 43 0.11 2.62 0.04 0.83 0.02 0.12 51020 P2 43 1.03 51020 UlP 50 1.08 2.63 0.03 0.90 0.08 51020 UlP 50 11.4 –10–50 m1 160 11.8 2.66 0.06 0.77 00.17–10 –5 0 m1 160 12 –10–50 m3 242 12.3 12.7 2.66 00.4 00–5 0 m3 242 12.6 –10–50 m4 242 12.7 2.67 0.13 0.47 00.39–10 –5 0 m4 242 12.75 51020 m4.1 343.8 12.95 2.68 0.07 0.51 00.424580120 m4.1 343.8 13 10 25 40 m4.2 364.9 13.1 2.67 0.05 0.56 00.34305080 m4.2 364.9 13.12 30 50 80 m4.3 377.2 13.2 2.67 0.08 0.24 00.67305080 m4.3 377.2 13.22 30 50 80 m4.4 409.3 13.28 2.69 0.27 0.07 00.664575110 m4.4 409.3 13.3 30 50 80 MFS 470.1 13.4 2.69 0.24 0.16 00.6 45 75 110 m4.5 478.6 13.55 2.65 0.04 0.82 00.14102550 m4.5 478.6 13.65 45 75 110 m5 502.1 13.75 2.65 0.07 0.56 00.374575110 m5 502.1 14.65 40 50 80 onlap uncf 511.8 14.8 2.67 0.08 0.51 00.33505580 (continued )

GEOSPHERE | Volume 12 | Number 5 Kominz et al. | Miocene relative sea level: IODP 313 New Jersey Margin 8 Research Paper

TABLE 2. AGES, AVERAGE DENSITIES (RHO), AVERAGE LITHOLOGIES, AND WATER DEPTHS* USED AS INPUT FOR SEQUENCES FROM SITES M27, M28, AND M29 (continued ) Average Average Average Average Depth Age Rho clay sand Csand silt WD lowWD mid WD high M29 input (continued ) downlap uncf 511.8 14.85 50 80 90 MFS 519.6 15.2 2.65 0.08 0.59 00.327590100 m5.2 602.2 15.5 2.67 0.18 0.27 0.01 0.54 70 75 100 m5.2 602.2 15.75 45 80 120 MFS 620.1 15.8 2.66 0.06 0.76 00.184580120 m5.3 643.2 16 2.68 0.12 0.48 00.37607080 m5.34 643.2 17.6 65 75 100 m5.4 662.4 17.7 2.68 0.19 0.19 00.576575100 m5.4 662.4 17.75 45 80 120 m5.45 683.2 17.85 2.68 0.16 0.34 00.474580120 m5.45 683.2 17.9 45 80 120 m5.47 687.9 18 2.64 0.14 0.31 00.414580120 m5.47 687.9 18.1 45 80 120 m5.6 710 18.3 2.67 0.14 0.38 0.01 0.42 45 80 120 m5.6 710 18.85 45 80 120 m5.7 728.6 19 2.68 0.14 0.40 0.42 45 80 120 m5.7 728.6 19.45 50 75 100 MFS 734.8 19.5 2.68 0.93 0.07 00 75 90 110 m5.8 746 19.65 2.69 0.88 0.05 00 45 75 85 m5.8 746 20.6 65 75 100 m6 755.5 20.9 2.64 0.16 0.36 00.486575100 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— strata, PH—undifferentiated and Holocene, P—Pleistocene and/or ; 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).

Stratigraphy above and below the Lower–Middle Miocene 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- Sediments beneath the Lower to Middle Miocene section sampled by Exp ogy and water depth of this younger part of the section have no impact on R1 313 have been compacted under the load of the Miocene and younger sedi- results. Therefore, the decompaction of the Lower–Middle Miocene focus of the ments. This compaction must be taken into account to properly estimate R1 paper is valid because the thickness of overlying strata is known and included for the Miocene strata. Based on a contour map of depth to basement (Ben- in the model. The thermal curve is fit to the R1 water-depth­ values based on the son, 1984), we estimate that ~8 km of pre-Miocene sediments were deposited pre-Miocene Anchor Dickinson sediments overlain by sites M27, M28, or M29 beneath the three offshore sites. For modeling purposes, we use the along- detailed stratigraphic data that are, in turn, overlain by 209, 243, or 342 m of strike Anchor Gas, Dickinson No. 1 rotary well (Poag, 1985; Olsson et al., 1988; younger strata observed at the latter three sites, respectively. The effects of the Sugarman et al., 2011; Fig. 1) to estimate the lithology and age data of the younger subsidence history on the magnitudes of Miocene relative sea level underlying strata. are minor, because the R2 results are registered to present sea level, which is The Lower to Middle Miocene strata also compacted beneath Upper Mio- unaffected by the intervening upper Miocene and younger section. Corehole cene and younger sediments. These sediments were either poorly sampled (Exp ­water depths at Sites M27, M28, and M29 are 34, 35, and 36 m below present

GEOSPHERE | Volume 12 | Number 5 Kominz et al. | Miocene relative sea level: IODP 313 New Jersey Margin 9 Research Paper

TABLE 3. AGES OF SURFACES FOR SHELF SEQUENCES AND AGES OF STRATA FOR ONSHORE SEQUENCES Sequence surface M27 age M27 JVB M28 ageM28 JVB M29 ageM29 JVB Kw ageOnshore sequence m4.1 12.96 13.2 ?13? m4.2 13 ?13.113.113.1 m4.2 13.1 13.1 13.1 13.1213.113.1 Kw3b m4.3 13.2 13.2 13.2 13.2 13.2 Kw3b m4.3 13.2 13.2213.2 Kw3b m4.4 ?13.28 13.3 Kw3b m4.4 13.4 ?13.36 ?13.313.3 Kw3b MFS 13.4 13.4 Kw3b m4.5 13.5 13.5 13.5513.313.55 13.6 13.5 Kw3b m4.5 13.65 13.6 13.5613.513.65 13.6 m5 13.75 13.7 13.7 13.7 13.7513.7 13.8 Kw3a 14.2 Kw3a m5 14.8 14.8 14.6514.814.65 14.6 onlap uncf 14.85 14.8514.8 downlap uncf 14.9 14.9 14.85 15 Kw2b MFS 15.2 15.2 15.2 Kw2b m5.2 15.3 15 15.4 15.1 15.5 15.6 Kw2b m5.2 15.6 15.6 15.7 15.7 15.7515.8 Kw2b MFS 15.8 15.8 15.8 Kw2b 15.9 Kw2b m5.3 15.95 15.8 16 16.3 16 16 16.1 Kw2a m5.3 16.65 16.5 16.5 16.6 Kw2a downlap uncf 16.55 Kw2a downlap uncf 16.6 Kw2a MFS 16.7 16.7 Kw2a m5.33 16.75 16.6 16.8 16.7 Kw2a m5.33 17.1 16.9 17.0517.4 Kw2a MFS 17.3 17.3 Kw2a m5.34 17.35 17 17.4517.617.5 Kw2a m5.34 17.65 17.7 17.6 17.6 17.6 17.6 MFS 17.67 17.67 m5.4 17.7 17.7 17.7 17.7 17.7 m5.4 17.79 ?17.79 17.9 17.7517.7 m5.45 17.85 18 17.851817.85 17.8 m5.45 17.86 17.9 17.9 17.9 17.9 Kw1c m5.47 17.9 18 18 18 Kw1c m5.47 18.1 ?18.118.1 Kw1c m5.6 18.3 18.3 18.3 18.3 Kw1c m5.6 18.6 18.8518.618.85 18.6 Kw1c m5.7 18.8 19 18.8 19 18.8 18.9 Kw1c m5.7 19.1 19.2 19.4 19.5 19.452019.2 Kw1b MFS 19.5 19.5 19.5 Kw1b Trans Surf 19.7 Kw1a m5.8 20.1 20.1 19.7 20 19.6520.2 Kw1a 20.4 Kw1a m5.8 20.7 20.7 20.5 20.5 20.6 20.5 m6 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.

GEOSPHERE | Volume 12 | Number 5 Kominz et al. | Miocene relative sea level: IODP 313 New Jersey Margin 10 Research Paper

sea level, respectively (Mountain et al., 2010). The registry to present sea level is well (Poag, 1985; Kominz et al., 1998). The beginning of thermal subsidence is a reasonable first assumption; although recent GIA effects on the order of 20 m taken as 140 Ma, the time when offshore sediment loading overcame lateral may complicate absolute sea-level estimates (Raymo et al., 2011) and are taken thermal uplift landward of the hinge zone (Steckler et al., 1988). For the Middle into account in our discussion in the Results section. 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- Coastal Plain Relative Sea-Level Curve tive sea level between deposition of these Miocene sediments and the present (Fig. 4B). 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 Combined Results, Coastal Plain Coreholes coastal plain relative sea-level estimates. However, during the two decades these drill core samples have been studied, revisions in the time scale—from Several coreholes drilled over the past decade in the New Jersey coastal Berggren et al. (1995) (BKSA95) to Lourens­ et al. (2004) and from the GTS2004 plain penetrate Lower and Middle Miocene strata, locally called the Kirkwood to the GTS2012 (Gradstein et al., 2012)—have resulted in small but significant Formation. These include coreholes at Ancora (Miller et al., 1999), Bass River adjustments to the Early to Middle Miocene. Early to Middle Miocene ages (Miller et al., 1998b), and Island Beach (Miller et al., 1994a), where only a few of are largely unchanged from BKSA95 (used for onshore sequences by previous the Kirkwood sequences were sampled and dated (Kw1a and Kw1b and possi- studies; e.g., Kominz et al., 2008) and offshore by Mountain et al. (2010) from bly Kw2a or b) and coreholes at Ocean View (Miller et al., 2001), Millville (Sugar- GTS2004 (used by Browning et al., 2013), except for the age of the / man et al., 2005), Cape May (Miller et al., 1996b), and Atlantic City (Miller et al., Miocene boundary. The Early to Middle Miocene time scale of GTS2004 dif- 1994b) that sampled multiple Kirkwood sequences. Early and Middle Miocene fers from GTS2012 only by placement of the / boundary R2 curves for these sites, in addition to the new Cape May Zoo site (Sugarman (0.16 m.y. older in GTS2012). All coreholes used in generating the coastal plain et al., 2007), show that several Kirkwood sequences are well represented (Fig. R2 curve have fairly well defined age ranges for the Miocene sequences, with 5A). In previous studies, the timing of sequences did not always match from age errors generally ±0.5 m.y. In sediments younger than 15 Ma, the errors can one corehole to the next due to uncertainties in assigning ages to sequences, be larger (i.e., ages are constrained by Sr isotopes that have an error of ±0.75 particularly in updip locations (e.g., the Ancora and Bass River sites). This re- to ±1.0 m.y.; Browning et al., 2013, their fig. 3). In downdip onshore coreholes, sulted in a blurring of the sequences so that sequence boundaries were not the age ranges are better constrained than they are at updip locations. We seen as hiatuses in the combined record (e.g., Kominz et al., 2008). While it is used the better dated sequences to constrain the less well dated coreholes. possible that the poor match is real, in most cases the uncertainty of age dates This generally resulted in age shifts of less than 0.5 m.y. and slightly greater is ~1 m.y. and larger in updip sites. We assume that the coastal plain sequence amplitudes in the sea-level estimates,­ because incidences of destructive inter- boundaries are regionally correlated, and we accept the ages of the best dated ference due to age miscorre­ lations­ have been reduced. The Cape May Zoo site sequences and adjust the other coreholes accordingly. Seismic correlations on- (Sugarman et al., 2007) was analyzed subsequent to the Kominz et al. (2008) shore and offshore confirm the correlations (Monteverde et al., 2008; Iscimen, compilation, and it has been included in our backstripped data set (see Re- 2014). In general, the Cape May Zoo sequences are well dated and provide ages sults section). Lithologies­ at all coastal plain sites were decompacted using the for the base of most sequences. However, for Kw1 and Kw2a, the sequences at

TABLE S3B: RELATIVE SEA LEVEL (RSL) AVERAGED NEW JERSEY COASTAL PLAIN. ALL DATA IS IN METERS Observed Observed Observed Shi ed Shi ed Shi ed coastal plain porosity versus depth relations of Van Sickel et al. (2004). Ocean View provide the best age constraints. Sequences Kw0, Kw1c, and Kw3a Age (Ma) RSL NJCP low RSL NJCP bestRSL NJCP high RSL NJCP low RSL NJCP bestRSL NJCP high 20.4 -33.414 -6.358 20.699 6.586 33.643 60.699 are best represented at the Cape May corehole. The tuned sequence ages allow 20.3 -3.395 9.911 23.217 36.438 49.744 63.051 20.2 -1.445 8.776 18.997 38.222 48.443 58.664 us to stack sequences from several coreholes; the composite distinguishes indi­ 20.1 -5.600 4.588 14.777 33.900 44.088 54.277 20.0 -3.170 3.693 10.557 36.163 43.027 49.890 RESULTS 19.9 -6.199 -3.884 -1.569 32.967 35.282 37.598 vidual sequences in the relative sea-level record (Fig. 5B). 19.8 -15.018 -8.294 -1.569 23.982 30.706 37.431 19.7 -20.136 -18.967 -17.797 18.697 19.867 21.036 The composite relative sea-level curve (Fig. 5C; data are provided in Sup- 19.6 -11.070 -7.520 -3.971 27.597 31.146 34.695 Coastal Plain Relative Sea-Level Curve 19.5 -10.358 -7.719 -5.081 28.142 30.781 33.419 4 19.4 -14.388 -10.293 -6.199 23.946 28.040 32.135 plemental Table S3 , section A) was generated following the procedures of 19.3 -18.182 -12.477 -6.772 19.985 25.690 31.395 19.2 -22.676 -14.935 -7.194 15.324 23.065 30.806 Cape May Zoo R1 and R2 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 4Supplemental Table S3. Relative sea-level curve for New Jersey Shelf and relative sea-level curve for New The Cape May Zoo corehole (Sugarman et al., 2007) provides additional low-end sea-level estimate. The best estimate relative sea level is taken as Jersey Coastal Plain. Please visit http://​dx.doi​ .org​ /10​ ​ Lower to Middle Miocene data that were not included in previous backstrip- the average of the high- and low-end estimates. Note that despite age model .1130/GES01241​ ​.S4 or the full-text article on www​ ping studies (Kominz et al., 2008). Younger sequences at this site are poorly tuning, we have not been able to separate sequences Kw1a and Kw1b that .gsapubs.org​ to view Supplemental Table S3. dated, and they are not included in our R2 analysis. The Oligocene and Eocene show no discernable hiatus at Ocean View or other coreholes. Nonetheless, portion of the R1 curve (Fig. 4A) is based on sediment from the Cape May we have separated sequences Kw3c into two sequences and separated Kw3a Site (Miller and Snyder, 1997), with older units based on the Anchor Dickinson from Kw3b and Kw1b from Kw1c. Error ranges have been reduced in some

GEOSPHERE | Volume 12 | Number 5 Kominz et al. | Miocene relative sea level: IODP 313 New Jersey Margin 11 Research Paper

Early Cretaceous Paleo. Eocene Olig. Miocene PP 0 A 100 50 Figure 4. (A) Cape May Zoo backstripping B Cape May Zoo results. R1 curves based on Cape May Zoo: 200 R1 values (Equation 2) older than the Mio- 25 cene are taken from the nearby Anchor Dickinson well (Fig. 1) that was drilled to basement. Unconformities are indicated 300 by gaps in R1 curves. Three R1 curves 0 Kw3b Kw3d are generated based on high, low (thin Kw3c orange lines) and best estimate (thick Kw2a Kw2b red line) paleoenvironmental estimates. 400 Kw1b Kw1c Relative Sea Level (m) -25 Also shown is the R1 result when no 20 18 16 14 12 10 paleo­water depths are taken into account (green curved line labeled “R1-sed”). Each 500 Age (m.y.) of the four R1 curves is fit to a thermal plate model. These best-fit curves are the Depth (m) smooth, continuous curves of the same 600 color as the R1 curves (see text for discus- Thermal sion). (B) Cape May Zoo R2 relative sea- model level curves. The difference between R1 fits and thermal plate model curves corrected 700 for water loading (Equation 3). Only the B Cape May Zoo well-dated Cape May Zoo relative sea level (R2) results are shown. These are part of 800 R1-sed the Kirkwood (Kw) Formation, as indicated in the figure. PP—Pliocene–Pleistocene; Olig.—Oligocene; Paleo.—. 900 140 120 100 80 60 40 20 0 Age (m.y.)

sequences as a result of matching high and low relative sea-level curves, and between sites (Fig. 6). To compare these new offshore results with our relative the timing of the sequences has shifted slightly, mainly due to changes in the sea-level curve generated using coastal plain boreholes, we construct a rela- geologic time scale. tive sea-level curve based only on the offshore sites, following the method of BurdigalianSLanghian errvallian Kominz et al. (2008) described above. That is, the error ranges are obtained 75 Seismic Sequences M27 M28 M29 high porosity best estimate by taking the minimum high estimate and the maximum low estimate of R2. ) m5.8 m5.5 m5.4 m5.3 m5.2 low porosity m5.6 50 m5 m4 Exp 313 Offshore Sites M27, M28, and M29 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 25 A critical test of the efficacy of one-dimensional backstripping for esti­ taken to be valid (Fig. 7A; results are provided in Supplemental Table S3 (sec-

Relative Sea Level (m mating relative sea level is the degree to which the calculated magnitude tion B, [footnote 4]). The “best estimate” R2 curve is based on an average of 0 of R2 is consistent among boreholes. In most instances and within errors of best estimates for all coreholes that sample sediment of that age. Where R2 Seismic Sequence Boundaries m4.4 m5.5 m4.3 m5.8 m5.7 m5.6 m5.4 m5.3 m5.2 m5 m5.33 m4.5 m4.2 water-depth estimates, our study passes this test for the three Exp 313 drill results from multiple coreholes do not overlap, the best estimate average does -25 m5.34 m4.1 20.0 17.5 15.0 12.5 sites (Fig. 6). Although using low- and high-end porosity versus depth curves not always fall between the high and low ranges (e.g., m5.3 and m5.2; Fig. 7). Age (m.y.) increases ranges of uncertainty, the form of m.y.-scale relative sea-level vari- In these cases, the error range has been extended to include the best estimate 5Supplemental Figure S2. Relative sea-level (R2) ations remains unchanged (see Supplemental Fig. S25). Sequences younger averages (Fig. 7B). Despite these inconsistencies, there is an overall good age curves from Integrated Ocean Drilling Program Ex- than sequence boundary m4 (ca. 12.6 Ma) are included in the modeling, but agreement, with all three offshore coreholes showing similar R2 trends and pedition 313 coreholes. Please visit http://​dx.doi​ .org​ ​ /10​.1130/GES01241​ ​.S5 or the full-text article on www​ they are not the focus of this study (see Table 2). Poor age and paleoenviron- magnitudes (Figs. 6 and 7). This suggests that we are reconstructing relative .gsapubs.org​ to view Supplemental Figure S2. ment control for these younger sequences could explain the lack of agreement sea-level change at the New Jersey shallow shelf.

GEOSPHERE | Volume 12 | Number 5 Kominz et al. | Miocene relative sea level: IODP 313 New Jersey Margin 12 Research Paper

Burdigalian Langhian Serrvallian 50

25

0 A

Relative Sea Level (m) Kw1c Kw2aKw2b Kw3a - Kw3c

–25 Kw1a & Kw1b

50 Ancora Bass River Coastal plain Figure 5. (A) Relative sea-level (R2) curves Ocean View Cape May Millville from Miocene sequences in coastal plain coreholes. Each corehole is indicated by its Atlantic City Cape May Zoo color in the key between 5A and 5B. Mio- cene stages are indicated on the top. The Early/Middle Miocene boundary occurs 25 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 B less well-dated correlative sequences. 0 (C) Composite Miocene relative sea-level Kw3a Kw3b Kw3c curve from New Jersey coastal plain core- holes using data plotted in panel 5B (see Kw2b Relative Sea Level (m) Supplemental Table S3 (section B) for coastal plain sea-level data [footnote 4]). –25 Kw1a Kw1b Kw1c Kw2a

25

0 C Kw3a Kw3b Kw3c Kw1a

Kw1c Kw2aKw2b

–25 Kw1b Relative Sea Level (m)

20.0 17.5 15.0 12.5 Age (m.y.)

GEOSPHERE | Volume 12 | Number 5 Kominz et al. | Miocene relative sea level: IODP 313 New Jersey Margin 13 Research Paper

100 Burdigalian Langhian Serrvallian

Seismic Sequences m5.2 m5.8 75 M29 m5.7 M28 m5.5 m5.3 m4 m5.4 m5.6 m5 M27

Figure 6. Relative sea-level (R2) curves from Integrated Ocean Drilling Program 50 (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 25 level (0 m) to roughly 80 m above modern. The sequences are labeled above the R2 values. Miocene stages are indicated at

Relative Sea Level (m) the top of the graph. The Early/Middle Miocene boundary occurs between the Burdigalian and Langhian Stages. 0

Seismic Sequence Boundaries m4.3 m5.8 m5.7 m5.6 m5.4 m4.4 m5.33 m5.3 m5.2 m5 m4.5 –25 m4.2 20.0 17.5 15.0 12.5 Age (m.y.)

DISCUSSION 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 The relative sea-level curve generated from offshore data shows higher GIA effects (Fig. 8A). From a sequence stratigraphic approach, we would ex- magnitudes of sea-level fluctuations than the onshore data (Fig. 8A). Taking into pect coastal plain sequence deposition to represent the most proximal portion account the full error range, the largest long-term sea-level range allowable in of transgressive and/or highstand systems tracts. These might occur during the shelf data is 75 m (16.7–13 Ma) compared to a 55 m range in the onshore data slowly rising relative sea level that may be associated with non-deposition off- (20.4–20.3 Ma). On the other , the minimum allowable sea-level ranges of shore. Thus, coastal plain sequences would be present during the maximum the two data sets are 32 m in the offshore data (16.7–12.2 Ma) and a 17 m in the amplitude of relative sea level. Our initial comparison of onshore and offshore onshore data (19.2–13.4 Ma). These ranges are entirely compatible with those sites (onshore is shown as light gray units in Fig. 8A) shows that deposition observed in comparable time intervals at the Marion Plateau by John et al. (2011). on the coastal plain sometimes occurs during hiatuses in the corresponding They report four sequences with maximum sea-level changes of 65 m and offshore sedimentation sampled by Exp 313. Where this is the case, it implies minimum ranges of 26 m between 13.8 and 16.6 Ma. All estimates are lower that the sequence boundaries of offshore units indicate relative sea-level rise than the 100–120 m variations implied by the global compilation of Haq et al. coupled with offshore sediment starvation rather than the more commonly ac- (1987), but are compatible with magnitudes of 20–50 m suggested by Haq and cepted model of sea-level fall (e.g., Loutit et al., 1988). In most cases, onshore Al-Qahtani (2005) for the . and offshore sequences show significant temporal overlap, with the offshore

GEOSPHERE | Volume 12 | Number 5 Kominz et al. | Miocene relative sea level: IODP 313 New Jersey Margin 14 Research Paper

BurdigalianSLanghian errvallian

Lowest high A Seismic Sequences m5.2 Highest low m4 75 m5.8 Average best m5.7 m5.4 Range overlap m5.5 m5.3 m5.6 m5 M29 M28 50 M27

25

0

B 50 Relative Sea Level (m) Average R2 with error ranges 25

0 Seismic Sequence Boundaries m4.4 m5.4 m4.3 m5.8 m5.7 m5.6 m5.3 m5.2 m5 m4.5 m4.2 m5.34 m5.33 –25 m4.1 20.0 17.5 15.0 12.5 Age (m.y.)

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.

GEOSPHERE | Volume 12 | Number 5 Kominz et al. | Miocene relative sea level: IODP 313 New Jersey Margin 15 Research Paper

Burdigalian Langhian Serrvallian

75 A Average R2 Seismic Sequences with error m5.8 ranges m5.7 m5.6 m5 m5.5 m5.3 m5.2 m5.4 coastal plain 50

25

Kw3c Kw3a 0 Kw1a Kw2b Kw1b Kw1c Kw2a Kw3b Average R2 B with error ranges –25 adjusted Relative Sea Level (m) coastal plain Kw1a Kw1b Kw3a Kw3b 50 Kw3c Kw2a Kw2b Kw1c

25

0 Seismice ssm Sequence Boundaries m4.3 m5.8 m5.7 m5.6 m5.4 m4.4 m5.33 m5.3 m5.2 m5 m4.5 m4.2

20.0 17.5 15.0 12.5 Age (m.y.)

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.

GEOSPHERE | Volume 12 | Number 5 Kominz et al. | Miocene relative sea level: IODP 313 New Jersey Margin 16 Research Paper

generally more complete. There are similar trends in offshore and onshore rected relative sea-level curves for onshore and offshore New Jersey provide a data in overlapping intervals. In fact, sequences Kw3b and Kw3c are, at least in working model for eustatic changes for the Early to Middle Miocene. part, compatible with relative sea level obtained from analyzing offshore strata It is premature to draw broad conclusions about the implications of these (Fig. 8A). Thus, in most cases, both onshore and offshore sequence bound­ results for the relationship between sequences and systems tracts for se- aries represent sea-level falls. quence paradigms because of limitations in the backstripping technique and The R2 sea-level estimates obtained from the onshore sites (0 ± 20 m above because of uncertainties in correlating sequences from onshore to offshore. present sea level) are considerably lower than those indicated by the offshore Such detailed comparisons are hampered by the fact that onshore data are sites (40 ± 30 m above present sea level), particularly in the older Miocene units. compiled from seven widely dispersed locations, while the offshore data set This is in direct contrast with stratigraphic principles that require sea-level rise consists of three coreholes along a single dip transect. to be higher when proximal land is flooded (e.g., Posamentier and Vail, 1988). Correlation of the data sets relies on both chronostratigraphy and the basic­ As such, it suggests that the New Jersey margin has experienced non-thermal assumptions of sequence stratigraphy (including superposition). All strata epeirogeny, resulting in differential subsidence of the onshore and offshore above a sequence boundary are younger than the strata below; this is also true portions of the margin. The dominant epeirogenic effect predicted to have oc- for maximum flooding surfaces. This allows correlation of sequences amongst curred in this region is the overriding of the Farallon plate. As the cold, dense onshore locations and offshore locations with finer resolution than provided lithosphere of the subducted slab passes beneath the margin, coastal New by the chronostratigraphy (e.g., we are reasonably certain that sequence Kw1a Jersey experiences subsidence (Conrad et al., 2004). Moucha et al. (2008) pre- is correlatable throughout the coastal plain coreholes). We are also reasonably certain of the onshore to offshore correlation not only through chronostratigra- dicted that over the past 30 m.y., the coastal plain of New Jersey has subsided phy but also through pattern matching and seismic correlations. For example, 50–100 m more than offshore New Jersey. The minimum value of 50 m of dif- the onshore composite sequence Kw2a sequence can be precisely correlated ferential subsidence over 30 m.y. implies a relative coastal plain uplift of 33 m with composite sequence m5.4 using well logs and seismic profiles (Iscimen, at 20 Ma, reducing to 20 m at 12 Ma. We add this differential subsidence to 2014). Thus, though the uncertainty on the numerical age of any given back- the onshore relative sea-level curve to make it comparable to the offshore R2 stripped estimate is large (±0.25 to ±0.5 m.y.), physical correlations (principles results. This correction for epeirogenic effects largely reconciles differences of superposition, seismic control, and sequence stratigraphy) mean that our between onshore and offshore relative sea-level estimates. relative age correlations amongst sites is finer. An additional variation between onshore and offshore relative sea level If deposition did not follow the tenets of sequence stratigraphy, then cor- is predicted by the effect of ice loading (GIA). Raymo et al. (2011) estimated relation would rely solely on chronostratigraphy. The age uncertainty on the the impact of ice volume and GIA effects on observations of Pliocene sea- ages of sequences is large (±0.25–0.5 m.y. offshore and ±0.5–1.0 m.y. onshore) level maxima. They found that GIA effects have diminished through time compared to the mean sea-level cycle duration of 0.84 ± 0.52 m.y. Exp 313 also for all but the and the very near-field location of the sampled several higher-order sequences (100/405 k.y.; Browning et al., 2013); melted ice sheets. However, the impact of the last glacial maximum on the thus the average duration of our sequences is shorter than 1 m.y. Never­the­ elevations of the New Jersey coastal plain and the New Jersey shelf estimated less, our sequences are of a higher order than predicted by changes in mantle by Raymo et al. (2011) will also affect older sea-level estimates by distorting dynamic topography that appear to operate on longer-term time scales (2– today’s datum.­ They found that the New Jersey coastal plain yields relative 100 m.y.; Petersen et al., 2010). Reliance on chronostratigraphy alone would sea-level estimates that are ~12 m above eustatic sea level, while the shelf not alter the pervasively lower relative magnitude of sea level recorded on yields relative sea-level estimates that are ~18–24 m above eustatic sea level the coastal plain as compared to that observed on the shelf. That is, relative (depending on distance offshore and on the assumptions made for lithosphere subsidence of the coastal plain would still be required, supporting the epeiro- rheology). Therefore, offshore relative sea-level estimates will always appear genic model. We expect more precise relative sea-level interpretations when to be 6–12 m higher than coastal plain estimates. Thus, to compare onshore the onshore and offshore core data are combined with seismic imaging in a relative sea-level maxima with offshore data, 6–12 m needs to be added to all full two-dimensional backstripping approach (e.g., Kominz and Pekar, 2001). onshore data. Rowley et al. (2013) also suggest five to ten meters of relative Future work will include additional correlations of offshore sequences using uplift occurred between the coastal plain and shelf as a result of GIA, con- seismic profiles and well logs (e.g., Iscimen, 2014) and two-dimensional back- sistent with the Raymo et al. (2011) results. The effect of the Farallon slab in stripping (e.g., Steckler et al., 1999). conjunction with the GIA effects shift relative sea-level variations predicted by In summary, comparison of R2 results from the new offshore coreholes the coastal plain curve relative to the offshore R2 curve (Fig. 8B). The resulting with the revised coastal plain sequences requires relative subsidence of the coastal plain adjusted-R2 magnitudes, in comparison with the offshore data, coastal plain since the Middle Miocene. This is compatible with modeling of are consistent with the assumption that the deposition on the onshore coastal mantle-driven subsidence due to the subducted Farallon slab (e.g., Moucha plain generally occurs at or near the highest magnitude of sea level. Our cor- et al., 2008).

GEOSPHERE | Volume 12 | Number 5 Kominz et al. | Miocene relative sea level: IODP 313 New Jersey Margin 17 Research Paper

CONCLUSIONS 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, Backstripped relative sea-level estimates of Middle to se- p. 227–248, doi:​10​.1111​/j​.1365​-2117​.2008​.00360​.x​. quences from the New Jersey coastal plain and three new IODP shelf sites are Browning, J.V., Miller, K.G., Sugarman, P.J., Barron, J., McCarthy, F.M.G., Kulhanek, D.K., Katz, generally internally consistent with respect to the timing of sequence bound- 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- aries and relative sea-level variations. Coastal plain sedimentation tends to sphere, v. 9 , p. 1434–1456, doi:​10​.1130​/GES00857​.1​. predict relative sea-level changes that are consistent with those seen offshore. Conrad, C., Lithgow-Bertelloni, C., and Louden, K.E., 2004, Iceland, the Farallon slab, and dy- Where onshore and offshore sedimentation corresponds, the rising and falling namic topography of the North Atlantic: Geology, v. 32, p. 177–180, doi:​10.1130​ /G20137​ ​.1​. portions of the relative sea-level curve can be correlated, although the magni- Grabau, A.W., 1936, Oscillation or pulsation: 16th International Geological Congress Report, v. 1, p. 539–553. tude of offshore and onshore estimates is offset. This result requires that the Gradstein, F.M., Ogg, J.G., Schmidtz, M.D., and Ogg, G.M., eds., 2012, The geologic time scale: New Jersey passive margin has undergone epeirogeny and, in particular, that New York, Elsevier, 1144 p. the offshore shelf has subsided less than the coastal plain. Thus, it is consistent Haq, B.U., and Al-Qahtani, A.M., 2005, cycles of sea-level changes on the Arabian Platform: Geoarabia, v. 10, p. 127–160. with relative epeirogeny due to subduction of the Farallon plate (Moucha et al., Haq, B.U., Hardenbol, J., and Vail, P.R., 1987, Chronology of fluctuating sea levels since the Tri- 2008) and due to GIA effects of the last deglaciation (Raymo et al., 2011). The assic (250 million years ago to Present): Science, v. 235, p. 1156–1167, doi:10​ .1126​ /science​ ​ amplitude of Late to Middle Miocene m.y.-scale sea-level changes seen at the .235​.4793​.1156​. Haq, B.U., Hardenbol, J., and Vail, P.R., 1988, and chronostratigraphy and New Jersey margin is generally 5–20 m but occasionally is as great as 50 m. 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​ ​ ACKNOWLEDGMENTS .88​.01​.0071​. Iscimen, T., 2014, Sequence stratigraphy of Miocene sequences Kw2a and m5.4, New Jersey: This manuscript was greatly improved by suggestions from one anonymous reviewer, reviewers Onshore to offshore correlations [M.S. thesis]: New Brunswick, New Jersey, Rutgers Uni- Hugo Pouderoux and Olivier Dauteuil, and Geosphere Associate Editor Jean-Noël Proust. This versity, 80 p. work was supported by funds from the Faculty Research and Creative Activities Award, Western John, C.M., Karner, G.D., Browning, E., Leckie, R.M., Mateo, Z., Carson, B., and Lowery, C., 2011, Michigan University to Kominz. Support was provided from National Science Foundation grants Timing and magnitude of Miocene eustasy derived from the mixed siliciclastic-carbonate EAR-1052257, OCE-1154379, and OCE14-63759 to Miller. Funding was supplied by the U.S. Sci- stratigraphic record of the northeastern Australian margin: Earth and Planetary Science Let- ence Support Program Consortium for Ocean Leadership to Katz, Miller, Browning, and Mountain. ters, v. 304, p. 455–467, doi:​10​.1016​/j​.epsl​.2011​.02​.013​. ­Samples were provided by the IODP. Katz, M.E., Miller, K.G., and Mountain, G.S., 2003, Biofacies and lithofacies evidence for paleo- environmental interpretations of upper sequences on the New Jersey continental shelf (ODP Leg 174A), in Olson, H.C., and Leckie, R.M., eds., Micropaleontologic Proxies for REFERENCES CITED Sea-Level Change and Stratigraphic Discontinuities: SEPM (Society for Sedimentary Geol- Ando, H., Oyama, M., and Nanayama, F., 2014, Data report: Grain size distribution of Miocene ogy) Special Publication 75, p. 131–146, doi:​10​.2110​/pec​.03​.75​.0131. successions, IODP Expedition 313 Sites M0027, M0028, and M0029, New Jersey shallow Katz, M.E., Browning, J.V., Miller, K.G., Monteverde, D., Mountain, G.S., and Williams, R.H., 2013, shelf, in Mountain, G., Proust, J.-N., McInroy, D., Cotterill, C., and the Expedition 313 Scien- Paleobathymetry and sequence stratigraphic interpretations from benthic : In- tists, Proceedings of the Integrated Ocean Drilling Program 313: Tokyo, Integrated Ocean sights on New Jersey shelf architecture, IODP Expedition 313: Geosphere, v. 9, p. 1488–1513, Drilling Program Management International, Inc., doi:​10​.2204​/iodp​.proc​.313​.201​.2014​. doi:​10​.1130​/GES00872​.1​. Benson, R.N., 1984, MS2 Structure contour map of Pre-Mesozoic basement, landward margin of Kominz, M.A., and Pekar, S.F., 2001, Oligocene eustasy from two-dimensional sequence strati- Baltimore Canyon Trough (North and South Sections): Delaware Geological Survey Miscel- graphic backstripping: Geological Society of America Bulletin, v. 113, p. 291–304, doi:​10.1130​ ​ laneous Map 2, scale 1:500,000. /0016​-7606​(2001)113​<0291:​OEFTDS>2​.0​.CO;2​. Berggren, W.A., Kent, D.V., Swisher, C.C., III, and Aubry, M.-P., 1995, A revised Cenozoic geochro- Kominz, M.A., Miller, K.G., and Browning, J.V., 1998, Long-term and short-term global Ceno- nology and chronostratigraphy, in Berggren, W.A., Kent, D.V., Aubry, M.-P., and Harde, J., zoic sea-level estimates: Geology, v. 26, p. 311–314, doi:​10​.1130/0091​ ​-7613(1998)026​ <0311:​ ​ eds., Geochronology, Time Scales, and Global Stratigraphic Correlations: SEPM (Society for LTASTG>2​.3​.CO;2​. Sedimentary Geology) Special Publication 54, p. 129–212, doi:​10​.2110​/pec​.95​.04​.0129​. Kominz, M.A., Browning, J.V., Miller, K.G., Sugarman, P.J., Misintzeva, S., and Scotese, C.R., Bond, G.C., and Kominz, M.A., 1984, Construction of tectonic subsidence curves for the early 2008, Late Cretaceous to Miocene sea-level estimates from the New Jersey and Delaware miogeocline, southern Canadian Rocky Mountains: Implications for subsidence coastal plain coreholes: An error analysis: Basin Research, v. 20, p. 211–226, doi:​10.1111​ ​/j​ mechanisms, age of breakup, and crustal thinning: Geological Society of America Bulletin, .1365​-2117​.2008​.00354​.x​. v. 95, p. 155–173, doi:​10​.1130​/0016​-7606​(1984)95​<155:​COTSCF>2​.0​.CO;2​. Kominz, M.A., Patterson, K., and Odette, D., 2011, Lithology dependence of porosity in slope and Bond, G.C., Kominz, M.A., and Grotzinger, J.P., 1988, Cambro- eustasy: Evidence from deep marine sediments: Journal of Sedimentary Research, v. 81, p. 730–742, doi:10​ .2110​ /jsr​ ​ geophysical modeling of subsidence in Cordilleran and Appalachian passive margins, in .2011​.60​. Paola, C., and Kleinspehn, K., eds., New Perspectives in Basin Analysis: New York, Springer-­ Kopp, R.E., Mitrovica, J.X., Griffies, S.M., Yin, J., Hay, C.C., and Stouffer, R.J., 2010, The impact of Verlag, p. 129–160, doi:​10​.1007​/978​-1​-4612​-3788​-4_7​. melt on local sea levels: A partially coupled analysis of dynamic and static equi- Bond, G.C., Kominz, M.A., Steckler, M.S., and Grotzinger, J.P., 1989, Role of thermal subsidence, librium effects in idealized water-hosing experiments: Climatic Change, v. 103, p. 619–625, flexure and eustasy in the evolution of early Paleozoic passive-margin carbonate platforms, doi:​10​.1007​/s10584​-010​-9935​-1​. in Crevello, P.D.,Wilson, J.L., Sarg J.F., and Read, J.F., eds., Controls on Carbonate Platform Lourens, L., Hilgen, F., Shackleton, N.J., Laskar, J., and Wilson, D., 2004, The Neogene Period, and Basin Development, SEPM (Society for Sedimentary Geology) Special Publication, in Gradstein, F., Ogg, J., and Smith, A., eds., A Geologic Time Scale: Cambridge University v. 44, p. 39–61, doi:​10​.2110​/pec​.89​.44​.0039​. Press, p. 409–440.

GEOSPHERE | Volume 12 | Number 5 Kominz et al. | Miocene relative sea level: IODP 313 New Jersey Margin 18 Research Paper

Loutit, T.S., Hardenbol, J., Vail, P.R., and Baum, G.R., 1988, Condensed section: The key to age Miller, K.G., Mountain, G.S., Wright, J.D., and Browning, J.V., 2011, A 180-million-year record of determination and correlation of continental margin sequences, in Wilgus, C.K., Hastings, sea level and ice volume variations from continental margin and deep-sea isotopic records: B.S., Kendall, C.G.St.C., Posamentier, H.W., Ross, C.A., Van Wagoner J.C., eds., Sea-Level Washington, D.C., Oceanography, v. 24, p. 40–53, doi:10​ .5670​ ​/oceanog​.2011​.26​. Changes: An Integrated Approach: SEPM (Society for Sedimentary Geology) Special Publi- Miller, K.G., Browning, J.V., Mountain, G.S., Bassetti, M.A., Monteverde, D., Katz, M.E., Inwood, cation 42, p. 183–213, doi:​10​.2110​/pec​.88​.01​.0183​. J., Lofi, J., and Proust, J.-N., 2013a, Sequence boundaries are impedance contrasts: Core- McCarthy, F.M.G., Katz, M.E., Kotthoff, U., Drljepan, M., Zanatta, R., Williams, R.H., Browning, seismic-log integration of Oligocene-Miocene sequences, New Jersey shallow shelf: Geo- J.V., Hesselbo, S.P., Bjerrum, C., Miller, K.G., and Mountain, G.S., 2013, Eustatic control of sphere, v. 9, p. 1257–1285, doi:​10​.1130​/GES00858​.1​. New Jersey margin architecture: Palynological evidence from IODP Expedition 313: Geo- Miller, K.G., Mountain, G.S., Browning, J.V., Katz, M.E., Monteverde, D., Sugarman, P.J., Ando, sphere, v. 9, p. 1457–1487, doi:​10​.1130​/GES00853​.1​. H., Bassetti, M.A., Bjerrum, C.J., Hodgson, D., Hesselbo, S., Karakaya, S., Proust, J.-N., and McKenzie, D., 1978, Some remarks on the development of sedimentary basins: Earth and Plane- Rabineau, M., 2013b, Testing sequence stratigraphic models by drilling Miocene foresets tary Science Letters, v. 40, p. 25–32, doi:​10​.1016​/0012​-821X​(78)90071​-7​. on the New Jersey shallow shelf: Geosphere, v. 9, p. 1236–1256, doi:10​ .1130​ /GES00884​ ​.1​. Miller, K.G., and Snyder, S.W., eds., 1997, Proceedings of the Ocean Drilling Program, Scientific Milne, G.A., Gehrels, W.R., Hughes, C.W., and Tamisiea, M.E., 2009, Identifying the causes of sea- Results, Leg 150X: College Station, , Ocean Drilling Program, 151 p. level change: Nature Geoscience, v. 2, p. 471–478, doi:10​ .1038​ /ngeo544​ ​. Miller, K.G., Wright, J.D., and Fairbanks, R.G., 1991, Unlocking the ice house: Oligocene-Mio- Monteverde, D.H., Mountain, G.S., and Miller, K.G., 2008, sequence development cene oxygen isotopes, eustasy and margin erosion: Journal of Geophysical Research, v. 96, across the New Jersey margin: Basin Research, v. 20, p. 249–267, doi:​10.1111​ ​/j​.1365-2117​ ​.2008​ p. 6829–6848, doi:​10​.1029​/90JB02015​. .00351​.x​. Miller, K.G., Browning, J.V., Liu, C., Sugarman, P., Kent, D.V., and Van Fossen, M., Queen, D., Moucha, R., Forte, A.M., Mitrovica, J.X., Rowley, D.B., Quéré, S., Simmons, N.A., and Grand, S.P., Goss, M., Gwynn, D., Mullikin, L., Feigenson, M.D., Aubry, M.-P., and Burckle, L.D., 1994a, 2008, Dynamic topography and long-term sea-level variations: There is no such thing as a Atlantic City site report, in Miller, K.G., et al., ed., Proceedings of the Ocean Drilling Program, stable continental platform: Earth and Planetary Science Letters, v. 271, p. 101–108, doi:10​ ​ Initial reports, Volume 150X: College Station, Texas, Ocean Drilling Program, p. 35–55. .1016​/j​.epsl​.2008​.03​.056​. Miller, K.G., Sugarman, P., Van Fossen, M., Liu, C., Browning, J.V., Queen, D., Aubry, M.-P., Mountain, G., Proust, J.-N., McInroy, D., and Cotterill, C., and the Expedition 313 Scientists, 2010, Burckle, L.D., Goss, M., and Bukry, D., 1994b, Island Beach site report, in Miller, K.G., et al., Proceedings of the Integrated Ocean Drilling Program, 313: Tokyo, Integrated Ocean Drilling eds., Proceedings of the Ocean Drilling Program, Initial Reports, Volume 150X: College Sta- Program Management International, Inc., 515 p., doi:​10​.2204​/iodp​.proc​.313​.101​.2010​. tion, Texas, Ocean Drilling Program, p. 5–33. Müller, D., Sdrolias, M., Gaina, C., Steinberger, B., and Heine, C., 2008, Long-term sea level fluc- Miller, K.G., Mountain, G.S., and Leg 150 Shipboard Party Members of the New Jersey Coastal tuations driven by ocean basin dynamics: Science, v. 319, p. 1357–1362, doi:​10.1126​ /science​ ​ Plain Drilling Project, 1996a, Drilling and dating New Jersey Oligocene–Miocene sequences: .1151540​. Ice volume, global sea level, and Exxon records: Science, v. 271, no. 5252, p. 1092–1095, doi:​ Olsson, R.K., Gibson, T.G., Hansen, H.J., and Owens, J.P., 1988, Geology of the northern Atlantic 10​.1126​/science​.271​.5252​.1092​. coastal plain: Long Island to Virginia, in Sheridan, R.E., and Grow, J.A., eds., The Atlantic Miller, K.G., Liu, C., Browning, J.V., Pekar, S.F., Sugarman, P.J., Van Fossen, M.C., Mullikin, L., Continental Margin, U.S.: Boulder, Colorado, Geological Society of America, Geology of Queen, D., Feigenson, M.D., Aubry, M.-P., Burckle, L.D., Powars, D., and Heibel, T., 1996b, Cape , v. I-2, p. 87–105. May site report, in Miller, K.G., et al., ed., Proceedings of the Ocean Drilling Program, Initial Parker, F.L., 1948, Foraminifera of the continental shelf from the Gulf of Maine to Maryland: Har- reports, Volume 150X (Supplement): College Station, Texas, Ocean Drilling Program, p. 1–28. vard Museum of Comparative Zoology Bulletin, v. 100, p. 213–241. Miller, K.G., Browning, J.V., Pekar, S.F., and Sugarman, P.J., 1997, Cenozoic evolution of the New Peltier, W.R., 1998, Postglacial variations in the level of the sea: implications for climate dynamics Jersey coastal plain: Changes in sea level, tectonics, and sediment supply, in Miller, K.G., and solid-Earth geophysics: Reviews of Geophysics, v. 36, p. 603–689, doi:​10.1029​ /98RG02638​ .​ and Snyder, S.W., eds., Proceedings of the Ocean Drilling Program, Scientific results, Vol- Petersen, K.D., Nielsen, S.B., Clausen, O.R., Stephenson, R., and Gerya, T., 2010, Small-scale ume 150X: College Station, Texas, Ocean Drilling Program, p. 361–373, doi:10​ .2973​ ​/odp​.proc​ mantle convection produces stratigraphic sequences in sedimentary basins: Science, v. 329, . sr​.150X​.326​.1997​. p. 827–830, doi:​10​.1126​/science​.1190115​. 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 Poag, C.W., 1981, Ecologic atlas of benthic foraminifera of the Gulf of Mexico: Stroudsburg, transect: Results from coastal plain and slope drilling: Reviews of Geophysics, v. 36, p. 569– Pennsylvania, Hutchison Ross Publishing Co., 174 p. 601, doi:​10​.1029​/98RG01624​. Poag, C.W., 1985, Depositional history and Stratigraphic reference section for central Baltimore Miller, K.G., Sugarman, P.J., Browning, J.V., Olsson, R.K., Pekar, S.F., Reilly, T.R., Cramer, B.S., Canyon Trough, in Poag, C.W., ed., Geologic Evolution of the United States Atlantic Margin: and Aubry, M.-P., Lawrence, R.P., Curran, J., Stewart, M., Metzger, J.M., Uptegrove, J., Bukry, New York, Van Nostrand Reinhold, p. 217–263. D., Burckle, L.H., Wright, J.D., Feigenson, M.D., Brenner, G.J., and Dalton, R.F., 1998b, Bass Posamentier, H.W., and Vail, P.R., 1988, Eustatic controls on clastic deposition II—Sequence and sys- River Site Report, Proceedings of the ODP, Initial Reports, 174AX: College Station, Texas, tems tract models, in Wilgus, C.K., Hastings, B.S., Kendall, C.G.St.C., Posamentier, H.W., Ross, Ocean Drilling Program, 39 p. C.A., and Van Wagoner, J.C., eds., Sea-Level Changes: An Integrated Approach: SEPM (Society Miller, K.G., Sugarman, P.J., Browning, J.V., Cramer, B.S., Olsson, R.K., de Romero, L., Aubry, for Sedimentary Geology) Special Publication 42, p. 125–154, doi:10​ .2110​ /pec​ .88​ .01​ ​.0125.​ M.‑P., Pekar, S.F., Georgescu, M.D., Metzger, K.T., Monteverde, D.H., Skinner, E.S., Uptegrove,­ Posamentier, H.W., Jervey, M.Y., and Vail, P.R., 1988, Eustatic controls on clastic deposition I—­ J., Mullikin, L.G., Muller, F.L., Feigenson, M.D., Reilly, T.J., Brenner, G.J., and Queen, D., Sequence and systems tract models, in Wilgus, C.K., Hastings, B.S., Kendall, C.G.St.C., Posa- 1999, Ancora Site, in Miller, K.G., Sugarman, P.J., Browning, J.V., et al., eds., Proceedings of mentier, H.W., Ross, C.A., and Van Wagoner, J.C., eds., Sea-Level Changes: An Integrated the Ocean Drilling Program, Initial Reports, Volume 174AX (Supplement): College Station, Approach: SEPM (Society for Sedimentary Geology) Special Publication 42, p. 109–124, doi:​ Texas, Ocean Drilling Program, p. 1–65. 10​.2110​/pec​.88​.01​.0109​. Miller, K.G., Sugarman, P.J., Browning, J.V., Pekar, S.F., Katz, M.E., Cramer, B.S., Monteverde, D., Raymo, M.E., Mitrovica, J.X., O’Leary, M.J., DeConto, R.M., and Hearty, P.J., 2011, Departures Uptegrove, J., McLaughlin, P.P., Jr., Baxter, S.J., Aubry, M.-P., Olsson, R.K., Van Sickel, B., from eustasy in Pliocene sea-level records: Nature Geoscience, v. 4, p. 328–332, doi:10​ .1038​ ​ Metzger, K., Feigenson, M.D., Tiffin, S., and McCarthy, F., 2001, Ocean View Site, in Miller, /ngeo1118​. K.G., Sugarman, P.J., Browning, J.V., et al., eds., Proceedings of the Ocean Drilling Program, Rowley, D.B., Forte, A.M., Moucha, R., Mitrovica, J.X., Simmons, N.A., and Grand, S.P., 2013, Initial Reports, Volume 174AX (Supplement): College Station, Texas, Ocean Drilling Pro- Dynamic topography change of the eastern United States since 3 million years ago: Science, gram, p. 1–72. v. 340, p. 1560–1563, doi:​10​.1126​/science​.1229180​. Miller, K.G., Kominz, M.A., Browning, J.V., Wright, J.D., Mountain, G.S., Katz, M.E., Sugarman, Royden, L., and Keen, C.E., 1980, Rifting process and thermal evolution of the continental margin P.J., Cramer, B.S., Christie-Blick, N., and Pekar, S.F., 2005, The Phanerozoic record of sea- of Eastern Canada determined from subsidence curves: Earth and Planetary Science Letters, level change: Science, v. 310, p. 1293–1298, doi:​10.1126​ /science​ ​.1116412​. v. 51, p. 343–361, doi:​10​.1016​/0012​-821X​(80)90216​-2​.

GEOSPHERE | Volume 12 | Number 5 Kominz et al. | Miocene relative sea level: IODP 313 New Jersey Margin 19 Research Paper

Spasojević, S., Liu, L., Gurnis, M., and Müller, R.D., 2008, The case for dynamic subsidence of man, P.J., Browning, J.V., et al., eds., Proceedings of the Ocean Drilling Program, Initial reports, the United States East Coast since the Eocene: Geophysical Research Letters, v. 35, L08305, Volume 174AX (Supplement): College Station, Texas, Ocean Drilling Program, p. 1–94. doi:​10​.1029​/2008GL033511​. Sugarman, P.J., Miller, K.G., Browning, J.V., Monteverde, D.H., Uptegrove, J., McLaughlin, J.P.P., Steckler, M.S., and Watts, A.B., 1978, Subsidence of the Atlantic-type continental margin off New Stanley, A.M., Wehmiller, J., Kulpecz, A., Harris, A., Pusz, A., and Kahn, A., 2007, Cape May York: Earth and Planetary Science Letters, v. 41, p. 1–13, doi:​10.1016​ /0012​ ​-821X​(78)90036​-5​. Zoo Site, in Miller, K.G., Sugarman, P.J., Browning, J.V., et al., eds., Proceedings of the Ocean Steckler, M.S., Watts, A.B., and Thorne, J.A., 1988, Subsidence and basin modeling at the U.S. Drilling Program, Initial reports, Volume 174AX (Supplement): College Station, Texas, p. 1–66. Atlantic passive margin, in Sheridan, R.E., and Grow, J.A., eds., The Atlantic Continental Sugarman, P.J., Monteverde, D.H., Pristas, R., Girard, M., Boyle, J., Miller, K.G., Browning, J.V., Margin: Boulder, Colorado, Geological Society of America, , v. I-2, Fan Reinfelder, Y., Romero, P., and Kulpecz, A., 2011, Characterization of the

p. 399–416. storage potential beneath the New Jersey Coastal Plain: Preliminary Characterization of CO2 Steckler, M.S., Mountain, G.S., Miller, K.G., and Christie-Blick, N., 1999, Reconstruction of Ter- Sequestration Potential in New Jersey and the Offshore Coastal Region, p. 1–44, http://​www​ tiary progradation and clinoform development on the New Jersey passive margin by 2-D .mrcsp​.org​/userdata​/phase​_ii​_reports​/njgs​_carbon​_sequestration​_report​_web​.pdf. backstripping: Marine Geology, v. 154, p. 399–420, doi:​10​.1016​/S0025​-3227​(98)00126​-1​. van Morkhoven, F.P.C.M., Berggren, W.A., Edwards, A.S., 1986, Cenozoic cosmopolitan deep-­ Sugarman, P.J., Miller, K.G., Browning, J.V., McLaughlin, P.P., Jr., Brenner, G.J., Buttari, B., Cramer, water benthic foraminifera: Pau, France, Elf-Aquitaine, Memoire 11, 421 p. B.S., Harris, A., Hernandez, J., Katz, M.E., Lettini, B., Misintseva, S., Monteverde, D.H., Olsson, Van Sickel, W.A., Kominz, M.A., Miller, K.G., and Browning, J.V., 2004, Late Cretaceous and R.K., Patrick, L., Roman, E., Wojtko, M.J., Aubry, M.-P., Feigenson, M.D., Barron, J.A., Curtin, S., Ceno­zoic sea-level estimates: Backstripping analysis of borehole data, onshore New Jersey: Cobbs, G., Cobbs, G., III, Bukry, D., and Huffman, B., 2005, Millville Site, in Miller, K.G., Sugar- ­Basin Research, v. 16, p. 451–465, doi:​10​.1111​/j​.1365​-2117​.2004​.00242​.x​.

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