Gold EOG Authors Manuscript

Download as pdf or txt
Download as pdf or txt
You are on page 1of 22

Sea-Level Change in Geological Time

David P Gold, CGG Robertson, Llandudno, Wales


© 2019.

Introduction 1

F
Sea-Level Cycles 1
Geological History of Sea-Level Change 5
Determining Sea-Level Change in the Geological Record 7

OO
Sedimentary Non-Uniqueness 8
Reconstructing Sea-Level Change 11
Stable Isotopes 11
Biological Sea-Level Indicators 14
Carbonate Sediments 16
Summary of Determining Sea-Level in the Geological Record 18

PR
Discussion 18
References 19
Further Reading 21

Introduction

D
The world’s oceans cover an area representing over 70% of the Earth’s surface, with saline water accounting for around 97.5% of Earth’s
total volume of water (Fig. 1). Sea-level, more accurately mean sea-level, is often regarded as the midpoint between mean low tide and
mean high tide at a given point along the Earth’s coastlines, and is a geodetic datum from which topographic elevations and bathymetric
TE
depths are measured. Traditionally in geology, the importance of sea-level has much to do with its position above or below the margin of
the continental shelf, marking highstand or lowstand conditions, respectively. Continental shelves comprise approximately 8% of the entire
area covered by oceans and are some of the Earth’s most biodiverse ecosystems (Fig. 1). The remaining 92% of the area covered by the
world’s oceans comprise continental slope and deep-water, bathyal and abyssal, environments (Fig. 1).
The total volume of water on Earth is estimated at between 1.3 and 1.5 billion cubic kilometers and this has remained relatively constant
EC

for around 4 billion years. The reason for this is that the hydrological cycle on Earth effectively operates as a closed system governed by
the laws of thermodynamics. That is, while energy is exchanged between the Earth and outer space in the form of radiation, for example,
significant amounts of matter, such as liquid water, neither enters nor leaves the system. However, quantitatively small amounts of water,
relative to the total global volume, may enter the Earth’s oceans through introduction by comets or from deep inside the Earth’s mantle.
Water may also leave the Earth by rising up through the stratosphere where hydrogen bonds in water molecules are broken apart by solar
radiation and hydrogen and oxygen atoms are emitted to space. Although the total volume of the Earth’s water has remained relatively
RR

stable throughout geological time, the volume of water locked up as glaciers and ice caps has varied considerably. The majority of water
stored in ice caps and glaciers is fresh water, and accounts for over two-thirds of the 2.5% of Earth’s water that is not saline (Fig. 1). A small
proportion of ice caps comprise frozen sea water. The melting of ice caps and glaciers introduces significant amounts of additional liquid
water into ocean basins, considerably increasing sea-level. In addition, the volume of seawater increases with rising global temperatures,
a process known as “thermal expansion,” further contributing to sea-level rise. Conversely, during glacial periods when a large volume of
CO

water is stored as snow and ice, sea-level is drastically reduced. Therefore sea-level change may be a direct reflection of changes in the
mass balance of ice sheets and glaciers, and the thermal expansion of water in warming oceans (Barlow et al., 2014).
Plate tectonic processes such as the formation and break-up of continents, and subduction and spreading of the ocean floor, together
with the cooling and thermal subsidence of oceanic lithosphere, increase and decrease the volume of ocean basins. These exert additional
controls on sea-level change. Therefore sea-level variation is controlled by two principal factors; changes in the volume of water in the
oceans and changes in the volume of ocean basins. Both these processes have caused great seas and oceans, such as the mighty Tethys and
Panthalassa, to appear and disappear as a response to the rise and fall of sea-level. Some seas still exist today as remnants of their former
UN

glory, or have made way for present day oceans, including the Atlantic and Pacific. The understanding of sea-level change throughout
geological time is not only important for oil and gas exploration, but also for understanding the response of the Earth’s marine ecosystems
to climate change today (Kemp et al., 2015).

Sea-Level Cycles

Sea-level change occurs, in the broadest possible sense, as a series of transgressions and regressions (Fig. 2). Marine transgression occurs
when there is a net rise in sea-level and sediment supply remains constant, while regression occurs when sea-level falls under the same

Earth Systems and Environmental Sciences, Volume ■ https://doi.org/10.1016/B978-0-12-409548-9.11899-8 1


2 Sea-Level Change in Geological Time

F
OO
PR
D
TE
EC

Fig. 1 Water distribution on Earth. The Earth acts a closed system where very little water is introduced to the system by comets and from the
Earth’s interior, or escapes into space. Most of Earth’s water comprises saline seawater, with a relatively small proportion of water locked up in ice
sheets and glaciers. This small amount of water has a profound influence on sea-level rise and fall as ice sheets recede and expand with fluctuating
global temperatures.
RR

conditions. Sea-level change is described as eustatic when it occurs globally and synchronously. This predominantly occurs as a response
to the volume of water within ocean basins controlled by the expansion and retreat of glaciers and ice caps, as described previously, or
through a change in the volume of ocean basins themselves. The volume of ocean basins may vary due to plate tectonic processes such as
dynamic topography and the Wilson cycle, including increased rates of seafloor spreading which form broad mid-ocean ridge systems, or
CO

the emplacement of large igneous provinces. Both events cause the displacement of water and eustatic sea-level rise. In addition, the geoid
provides a mantle-derived control on global sea-level. The removal of old mid-ocean ridge systems through subduction, or the subsidence
of old, cold and dense oceanic crust, increases the volume of ocean basins, resulting in eustatic sea-level fall.
Isostatic sea-level change occurs where localized glacial or tectonic loading cause regional subsidence, leading to relative sea-level rise
(Fig. 3). While eustatic sea-level is measured from a reference point at the center of the Earth, relative sea-level is the position of mean
sea-level relative to the position of the crust (Fig. 3). Therefore, vertical movements of the crust through tectonic subsidence or uplift can
UN

produce localized effects of relative sea-level rise or fall (e.g., Rovere et al., 2016). For example, relative sea-level rise can occur in an
area where the rate of tectonic subsidence exceeds the rate of eustatic sea-level fall. The concept of ‘forced’ and “unforced” or “normal”
regressions is dependent on the concept of relative sea-level. Forced regression occurs when relative sea-level fall pushes the position of
the shoreline seaward, regardless of whether the rate of sediment supply is high. Normal regression may occur when the rate of sediment of
supply exceeds the rate at which relative sea-level is rising, causing the position of the shoreline to move seaward. Other reference frames
to determine sea-level change include the concept of base-level, an equilibrium surface between net erosion and net deposition. Base-level
is positioned slightly below mean sea-level due to the erosion of rivers, waves and marine currents which can occur below sea-level
(Catuneanu, 2006). One other concept is the topographic equilibrium surface which is produced where there is no tectonic subsidence or
uplift, the rate of sediment deposition is equal to the rate of erosion and environmental conditions remain the same (Thorne, 1992).
Sea-Level Change in Geological Time 3

F
OO
PR
Fig. 2 Diagram depicting different stages of various sea-level cycles. One complete cycle is recorded by a sine curve where oscillations represent
sea-level rise and fall. The most basic sea-level cycle is the transgressive-regressive cycle reflecting sea-level rise and fall, respectively. Sequence
stratigraphic definitions of sea-level change define three or four stages, or systems tracts, within the cycle including lowstand, transgressive,
highstand and occasionally falling-stage systems tracts.

D
TE
EC
RR
CO
UN

Fig. 3 Eustatic sea-level change is measured from a fixed reference point, often the center of the Earth. Eustatic sea-level change occurs slowly,
globally and synchronously often as a response to the expansion and retreat of ice sheets and glaciers at the poles. Relative sea-level change is
measured from a position on the Earth’s crust, such as the continental shelf, whereby relative sea-level rise may occur in response to local tectonic
subsidence.

The principles of sequence stratigraphy are intrinsically linked to cycles of sea-level, as systems tracts are interpreted to form under
different positions on a sea-level curve. The sea-level cycle is expanded by sequence stratigraphic interpretations into three or four
systems tracts which occur within one complete cycle of sea-level rise and fall (Fig. 2). Within this cycle, transgressive systems tracts
4 Sea-Level Change in Geological Time

(TSTs) occur on the rising, or transgressive, limb of the sea-level trend (Fig. 2). Following this transgressive phase, highstand systems
tracts (HSTs) occur either side of the maximum transgressive inflection of the sea-level cycle, which marks the point at which a maximum
flooding surface forms (Fig. 2). Sedimentary deposits and depositional environments typically interpreted to be associated with TSTs and
HSTs include estuaries, carbonate platforms and build-ups including, but not limited to, reefs susceptible to highstand shedding. In some
sequence stratigraphic models, falling stage systems tracts (FSSTs) occur on the falling, or regressive, limb of the sea-level trend (Fig.
2). However, many sequence stratigraphic models are tripartite and only incorporate transgressive, highstand and lowstand systems tracts
(LSTs). The FSST is therefore equivalent to the late HST and early LST (Fig. 2). LSTs occur either side of the maximum regressive

F
inflection of the sea-level cycle. At this stage, large parts, if not all, of the continental shelf are exposed and subaerial erosion processes
may form an unconformity and sequence boundary. Depositional sequences within sequence stratigraphic models are bounded at their top
and base by unconformities formed as sequence boundaries during periods of sea-level lowstand. Sedimentary deposits and depositional

OO
environments typically interpreted to be associated with FSSTs and LSTs include shelf edge deltas, turbidites and fluvially-incised shelves.
Sequence boundaries should not be recorded within deep water strata deposited below the margin of the continental shelf. However, where
a subaerial exposure surface is recorded in deposits from shallower water settings, this surface may be traced into deeper water settings as
a correlative conformity.
Changes in the volume of liquid water or volume of ocean basins may take many millions or only several thousands of years to occur.
These changes are often described in orders of magnitude (Table 1). Changes in the volume of ocean basins driven by plate tectonics

PR
often occur slowly, over periods of many tens of millions of years and are described as low, first or second order (Table 1; Fig. 4)

Table 1 Orders and magnitude of sea-level change with possible mechanisms.

Years (power of
Order Years (number) Mechanism
10)

D
1st 108 100,000,000 Supercontinent fragmentation and construction (Wilson cycle), opening and closing of
ocean basins
TE
2nd 107 10,000,000 Passive margin subsidence, major inter/glacial episodes
3rd 106 1,000,000 Global eustatic sea-level change, rifting and thermal subsidence
4th 105 100,000 Orbital forcing, minor inter/glacial episodes, regional tectonic events, sedimentary
5th 104 10,000 Orbital forcing, minor inter/glacial episodes, regional tectonic events, sedimentary

Modified from Tucker, M.E. and Wright, V.P. (1990). Carbonate sedimentology. John Wiley & Sons.
EC
RR
CO
UN

Fig. 4 Sea-level change may occur over a variety of different time scales, with durations lasting from several thousand to several hundred million
years. Short-term, high order, changes in sea-level may be superimposed on longer term, low order, trends.
Sea-Level Change in Geological Time 5

whereas changes in the volume of water that effect eustatic sea-level and often interpreted to be responsible for the deposition of systems
tracts, can occur relatively rapidly over a period of tens of thousands of years and are described as high, fourth or fifth order (Table 1;
Fig. 4). Globally recognized sequence boundaries and maximum flooding surfaces used in sequence stratigraphic interpretations (e.g.,
Hardenbol et al., 1998) correspond to third order cycles of global eustatic sea-level change (Table 1). Short-term, high order sea-level
change may be superimposed on longer term, low order trends (Fig. 4). Therefore in geological sequences a combination of high and low
order sea-level cycles may be recorded.
Milankovitch cycles represent the highest order of sea-level fluctuations, with oscillations occurring over time intervals of a few tens

F
of thousands of years. Milankovitch cycles are also known as orbital cycles as they are influenced by the spin of the Earth’s axis and its
orbit around the sun, driving short-term climate variation. These cycles are a function of three parameters, namely eccentricity, obliquity,
and precession. Eccentricity is controlled by changes in the shape of Earth’s orbit. The Earth’s orbit varies from nearly circular to elliptical,

OO
predominantly controlled by the gravitational pull of the planets Jupiter and Saturn, in cycles of approximately 100,000 years (Fig. 5).
When the Earth’s orbit is elliptical, the furthest distance from the Earth to the Sun is termed the aphelion while the closest distance is
termed the perihelion (Fig. 5). Measurements of orbits commonly range between 0, representing a circular orbit, and 1, forming an elliptical
orbit. Today, the Earth’s orbit is approximately 0.017, nearly circular, and is around 5 million kilometers closer to the Sun at perihelion in
January (147.5 million kilometers) than at aphelion in July (152.6 million kilometers). Solar radiation is approximately 7% less intense at
aphelion than at perihelion. However, temperatures are around 2.3°C warmer at aphelion as this occurs during the Northern Hemisphere

PR
summer. When the Earth’s orbit is at its most elliptical, however, solar radiation is approximately 23% more intense at perihelion compared
to aphelion and this may cause global temperatures and eustatic sea-level to rise as ice sheets recede.
Obliquity of the elliptic, or put simply obliquity, refers to the variation in the degree of tilt of the Earth’s axis (Fig. 5). The tilt of the
Earth’s axis varies between 22.1° and 24.5° (an angle of 2.4°) in cycles of approximately 41,000 years (Fig. 5). Today, the Earth is half
way through its obliquity cycle and has an axial tilt of 23.44°. When the angle of the Earth’s axial tilt is greatest (at 24.5°) there is more
variation in the amount of solar radiation delivered to the surface of the Earth between the seasons, and it is at its most intense during the
summer months of the Northern Hemisphere than when the Earth’s axial tilt is shallower (22.1°). Precession occurs in cycles of around

D
26,000 years and refers to the ‘wobble’ of the Earth’s axial rotation, similar to the wobble observed on a spinning-top. This motion occurs
due to the gravitational pull of the Sun and the Moon on the water at the Earth’s equatorial bulge, which also causes the shortest term
sea-level change in the form of tides. The equatorial bulge itself is formed by the centrifugal force exerted by the Earth’s axial rotation on
TE
its oceans.

Geological History of Sea-Level Change


EC

The history of sea-level change throughout Earth’s geological past can broadly be categorized into greenhouse and icehouse periods (Fig.
6). Major glacial episodes occurred within the Proterozoic which included at least two global ice ages during the Cryogenian, and the
Phanerozoic (Fig. 6) which included glaciations in the Ordovician to Silurian (Andean-Saharan), Carboniferous (Karoo) and the Late
Cenozoic incorporating today’s glacial period. During these major glacial episodes, much of the Earth’s water was locked in extensive ice
sheets resulting in periods of low sea-level with higher sea-level in the intervening greenhouse periods (Fig. 6). These greenhouse-icehouse
cycles are responsible for long duration, low order marine transgressions and regressions (Table 1).
RR

The Cryogenian marked an important stage in Earth history. Four major glaciations occurred during this time period, two of which, the
Marinoan and Sturtian, are interpreted to have occurred globally and were almost irreversible due to the albedo effect. The result of these
glaciations has become termed “Snowball Earth” and each major glacial period was followed by the deposition of continuous layers of
cap carbonates above Neoproterozoic glacial deposits or sub-glacial erosion surfaces. Deglaciation impacted ocean chemistry, which may
have contributed to the evolution of complex metazoans organisms during the Phanerozoic. Periods of major silicate weathering during
deglaciation may have caused a large flux of nutrients and metals into the ocean. This flux could have been greater than the present day and
CO

may have resulted in the oversaturation of the ocean with nutrients.


During the Phanerozoic, a combination of plate tectonics and alternating greenhouse and icehouse conditions drove the formation of
vast, and geologically important, seas and oceans. One of the first great oceans was the Iapetus, situated between the paleocontinents
of Laurentia, Baltica, and Avalonia in the Southern Hemisphere from the Ediacaran Period of the Late Neoproterozoic to the Devonian
Period of the Early Paleozoic. The supercontinent Gondwana was situated to the east of Laurentia, Baltica and Avalonia, and was
separated from these paleocontinents by the Rheic and Paleo-Tethys Oceans (Fig. 7). To the north of Gondwana, Laurentia, Baltica and
UN

Avalonia the Panthalassic Ocean covered much of the Northern Hemisphere. In the Iapetus, Rheic and Paleo-Tethys Oceans the first
complex, multicellular life forms evolved, including the Ediacaran biota, followed by the “Cambrian Explosion” in which shelled organisms
developed, followed by the first appearance of trilobites and echinoderms. From the Ordovician to Silurian, the first land plants and insects
colonized the land while jawed fish (gnathostomes) and the first sharks appeared in the world’s oceans. During this time, the first major
glaciation of the Phanerozoic, the Andean-Saharan glaciation, also occurred. Evidence of this glaciation is found in the form of glacial
dropstones, diamictites and moraines recorded in strata from Africa and South America (Le Heron et al., 2010), which were situated close
to the South Pole at this time (Fig. 7).
The Caledonian Orogeny during the Silurian resulted in the formation of the supercontinent of Laurussia, also known as Euramerica
(Fig. 7). This supercontinent was formed by the collision of the Laurentia, Baltica and Avalonia paleocontinents as they drifted north
from the Southern Hemisphere towards the equator (Stow, 2012). From the Devonian, Laurussia was separated from Gondwana to the
6 Sea-Level Change in Geological Time

F
OO
PR
D
TE
EC
RR
CO
UN
Sea-Level Change in Geological Time 7

southeast by the Rheic Ocean where lobe-finned fish (sarcopterygians) first began to crawl out of the water giving rise to the first land
animals. Plate convergence that caused the Caledonian Orogeny continued into the Devonian, resulting in the closure of the Rheic and
Iapetus Oceans (Stow, 2012) and leading to the collision of Laurussia and Gondwana to form the supercontinent of Pangea during the
Variscan, or Hercynian, Orogeny (Fig. 7). The closure of the Rheic and Iapetus Oceans and formation of Pangea in the Carboniferous
created a large landmass in the southern polar region and disrupted global warm water currents in the surrounding Panthalassic and
Paleo-Tethys Oceans. This triggered an expansion of the ice sheets, leading to the second major Phanerozoic glaciation known as the
“Karoo Ice Age” (Fig. 7).

F
The supercontinent Pangea had a large embayment to the east, comprising the remnants of the Paleo-Tethys Ocean, while on its northern
and western coastlines it was surrounded by the Panthalassic Ocean (Fig. 7). Pangea persisted from the Carboniferous until its final
break-up in the Jurassic. However, rifting initiated in the Late Permian to Early Triassic with the break off of the Cimmerian Superterrane.

OO
As the Cimmerian Superterrane drifted north, the Paleo-Tethys Ocean to the north of the terrane was subducted, and the Neo-Tethys
Ocean opened to the south (Fig. 7). By the Late Jurassic, the Cimmerian Superterrane had collided with Eurasia, resulting in the closure
of the Paleo-Tethys. The Neo-Tethys thus became the Tethys Ocean (Fig. 7). During the Late Paleozoic and Mesozoic, the Tethys and
Panthalassic Oceans were home to diverse marine ecosystems that included orthocones, ammonites, rudist bivalves and marine reptiles
such as ichthyosaurs, plesiosaurs, pliosaurs and mosasaurs. Strata deposited within the shallow water settings of the Tethys Ocean formed
important oil and gas reservoirs in the Middle East.

PR
The break-up of Pangaea occurred in stages as a series of rifting events. During the Early-Middle Jurassic, an east-west trending rift
separated north and south Pangea back into the continents of Laurasia and Gondwana, respectively, forming a seaway that linked the
western Tethys and eastern Panthalassa. This was followed by a series of north-south trending rifts which began splitting western from
eastern Pangea and opening up the new Atlantic Ocean. The opening of the Atlantic Ocean initiated in the north and propagated south,
leading to the separation of Africa from South America by the Early Cretaceous. During this time, the other Gondwanan fragments,
Antarctica, India and Australia, also began separating. Africa, India and Australia began to move north relative to Antarctica with India
moving at a faster rate than the other fragments.

D
During a greenhouse period in the Late Jurassic and Cretaceous there was a steady long-term increase in eustatic sea-level, reaching
between 200 m and 300 m higher than present day sea-level and resulting in 82% of the Earth’s surface being submerged under water at
its peak in the Late Cretaceous (Snedden and Liu, 2010; Stow, 2012). There were several mechanisms for the cause of this sea-level rise.
TE
Not only were ice caps absent from the poles due to high global temperatures, thermal expansion of seawater due to high temperatures
also contributed. Importantly, an equatorial east-west trending mid-ocean ridge that formed along the central axis of the Tethys Ocean
between Laurasia and Gondwana during the break-up of Pangaea created a large submarine mountain belt which displaced ocean waters
and contributed to global eustatic sea-level rise (Stow, 2012). This increase in global sea-level resulted in the flooding of many continental
interiors. The most famous of these include the Trans-Saharan Seaway of northern Africa, and Western Interior and Hudson Seaways of
EC

North America (Stow, 2012). The North American seaways separated the landmasses of Laramidia in the west from Appalachia to the east.
By the mid-Paleogene, India had collided with Asia, and the Atlantic and Pacific were the dominant oceans. Throughout the Neogene,
the remnants of the once mighty Tethys Ocean became the Mediterranean Sea which was gradually closed as Africa continued to move
northwards towards Europe. In addition, the continued northward movement of Australia, together with complex regional tectonics in
Southeast Asia, began separating the Indian and Pacific Oceans. Today these oceans are only connected by the highly tortuous “Indonesian
Throughflow.” During the Pliocene the uplift of the Isthmus of Panama connected North and South America, causing a great faunal
RR

interchange of land mammals and the creation of the Gulf Stream and North Atlantic Current which deliver warm waters from the Gulf
of Mexico to Northwest Europe. The third major Phanerozoic glaciation began in the Pleistocene Epoch. This epoch comprised a series of
glacial and interglacial periods driven by Milankovitch Cycles, causing high order sea-level oscillations. Glacial melting induced sea-level
rise occurred during the current interglacial, or Holocene period from 12,000 years ago until Present. However, the rate of melting and
proportional sea-level rise was not constant. The rate of glacio-eustatic sea-level rise was incredibly rapid during warmer periods of the
Holocene, but ceased or fell during cooler periods. The rapid rate of sea-level rise of up to 10 m per 1000 years during this time (Donovan
CO

and Jones, 1979), was known as the “Flandrian Transgression.” This was enough to outpace carbonate production rates (Table 2) leading
to carbonate platform drowning in sub-tropical belts, including the Arabian Gulf.

Determining Sea-Level Change in the Geological Record


UN

The Swiss geologist Eduard Seuss was the first to link global changes in sea-level to sedimentary sequences in the geological record,
defining the concept of eustasy (Rovere et al., 2016). There are many examples of global eustatic sea-level curves created for the
Phanerozoic (e.g., Haq and Al-Qahtani, 2005; Miller et al., 2005; Kominz et al., 2008; Snedden and Liu, 2010). However, much
recent work has demonstrated that many of the sedimentary deposits traditionally associated with the different systems tracts of the
sequence stratigraphic approach (see “Sea-Level Cycles” section) are in fact not unique and can form at any point on the sea level cycle

Fig. 5 Milankovitch, or orbital, cycles controlled by eccentricity, obliquity and precession of the Earth as it orbits around the sun.
8 Sea-Level Change in Geological Time

F
OO
PR
D
TE
Fig. 6 Icehouse and greenhouse periods during the Phanerozoic compared with eustatic sea-level curve of Snedden and Liu (2010). The major
glaciations correspond to short- and long-term falls in global sea-level. Conversely, periods of high sea-level occur during greenhouse periods.

under the right conditions. Sedimentary sequences that are shown to have been deposited under the control of eustatic sea-level cycles
EC

typically display laterally continuous “layer cake” facies changes (Burgess, 2006). However, most sedimentary sequences are much more
complex and exhibit lateral heterogeneity with complex stacking patterns and variable thicknesses, interpreted to be controlled by a variety
of different factors including rates of sediment supply or production, sediment transport and erosion, and local environmental or climatic
factors (Burgess, 2006; Burgess and Wright, 2003). Recognizing sea-level change in the geological record therefore requires identifying
unequivocal evidence for changes in water depth.
RR

Sedimentary Non-Uniqueness
Growing evidence for the non-uniqueness of stratal geometries, particularly through the use of numerical stratigraphic forward models, is
making sequence stratigraphic interpretations of sea-level change largely redundant. A good example of the non-uniqueness of sedimentary
sequences is the argument on the formation of shelf-edge deltas. Early sequence stratigraphic interpretations assumed that terrigenous
CO

sediment delivery to the deep ocean occurs during relative sea-level fall, whereby shoreline regression pushes the position of prograding
deltas to the shelf edge (Zhang et al., 2017). However, more recent numerical stratigraphic forward modeling and field studies show that
shelf-edge deltas can occur under conditions of rising sea-level (highstand), as well as lowstand conditions (Zhang et al., 2017). Together
with sediment supply, shelf width is a key control on the formation of shelf-edge deltas during rising sea-level (Burgess et al., 2008; Zhang
et al., 2017).
Numerical stratigraphic forward modeling results of, Burgess and Hovius (1998), Burgess et al. (2008) and Zhang et al. (2017)
UN

reflect observations of modern deltaic systems (e.g., Walsh and Nittrouer, 2003; Sweet and Blum, 2016) in that narrow, low-gradient
shelves, and the deltas that form on them under conditions of long-lived low-amplitude eustatic sea-level rise and slow shelf subsidence,
favor delta formation reaching the shelf edge. In addition, global climate may also play an important role in the formation of shelf-edge
deltas. In numerical forward models of 32 modern shelves formed under different eustatic rise rate and duration scenarios, 28% of shelves
formed highstand deltas under icehouse conditions while 97% formed highstand deltas during greenhouse conditions (Zhang et al., 2017).
Combined with analyses of shelf width during icehouse conditions, 50% of shelves < 50 km in width produced highstand shelf-edge deltas
whereas only 18% of shelves > 50 km produced highstand shelf-edge deltas (Zhang et al., 2017).
Eustatic sea-level change and shelf subsidence have been shown to play a less important role in the formation of shelf-edge deltas
(Zhang et al., 2017). These conclusions challenge traditional sequence stratigraphic assumptions (e.g., Catuneanu, 2006; Catuneanu et
al., 2009) that suggest that the major controls on the deposition of sedimentary sequences are limited to sediment supply, shelf subsidence
Sea-Level Change in Geological Time 9

F
OO
PR
D
TE
EC
RR
CO

Fig. 7 Paleogeographic reconstructions of major continents and oceans throughout Earth history from the Cambrian to Pleistocene. The major
glaciations occur within the Silurian (Andean-Saharan), Carboniferous (Karoo) and Late Cenozoic (Recent glaciation). During the Cretaceous
greenhouse period, high sea-level resulted in the flooding of many continental interiors creating seaways including the Western Interior and
Trans-Saharan. The Panthalassic Ocean occupied approximately 70% of the Earth’s surface from the Paleozoic to Early Mesozoic before the
UN

development of the Pacific Ocean. The Tethys Ocean closed and disappeared by the end of the Miocene as the African continent moved north
towards Europe.

and eustatic sea-level change, rather than morphology (Zhang et al., 2017). There are many more examples of where traditional sequence
stratigraphic concepts do not adequately describe the deposition of sedimentary sequences.
Examples from modern and ancient systems (e.g., Carvajal and Steel, 2006; Covault and Graham, 2010; Castelltort et al., 2017)
demonstrate that submarine fans can also form under rising sea-level, contrary to their formation within lowstand wedges of traditional
approaches (Zhang et al., 2017). The rate of relative sea-level fall is commonly described to control the delivery of sand to deep water
areas, as a reduction in accommodation space on the shelf causes sedimentation to shift basinward (Harris et al., 2018). However,
10 Sea-Level Change in Geological Time

Table 2 Typical carbonate production rates.

Environment Carbonate production rates (mm yr-1/m Ka− 1/km Ma− 1)

Tropical carbonate production 0.2–10 (usually 0.5–1)


Benthic foraminifera, corals and calcareous algae on seaward reef flats 0.5–1.5
Back-reef lagoons 0.1–0.5
Reef front (coral growth) Up to 6

F
Active corallites 50–100
Temperate carbonate production 0.01–0.2
Passive margin subsidence 0.01 (sag)-2.5 (rift)

OO
3rd order plate tectonic sea-level rise 0.01
4th order glacio-eustatic sea-level rise 0.1–10

After Tucker, M.E. and Wright, V.P. (1990). Carbonate sedimentology. John Wiley & Sons.

PR
numerical stratigraphic forward modeling results of Harris et al. (2016, 2018) show that there is a poor relationship between deep-water
sand delivery and sea-level, and that deep water sand delivery can occur at any stage on the sea-level cycle at different rates of relative
sea-level change. Linear regression analyses of numerical stratigraphic forward models that test the relationship between sea-level change
and deep-water sand delivery by varying the values of water-driven transport of sand over several orders of magnitude show the rate
of relative sea-level change to be a less important factor in controlling sand delivery to the deep ocean, accounting for only 1% of the
variability of the deep-water sand delivery rates regardless of transport efficiency (Harris et al., 2018). The position of relative sea-level
in the sea-level cycle itself accounts for up to 55% of the variability as it is the occurrence of a sea-level fall that promotes sand delivery

D
to deep water better than the rate of fall (Harris et al., 2018). The remaining 45% of the variability in deep-water sand delivery rates is
attributed to oceanographic or morphological characteristics such as climate, high sediment transport rates and shelf width, or stochastic
(random) processes such as seismic activity, slope failure or the abandonment and formation of new channels (Harris et al., 2018).
TE
Within the sequence stratigraphic model, the formation of sequence boundaries is attributed to sea-level fall causing valleys cut
by fluvial systems to incise the shelf (Howell et al., 2018). One of the main testing areas of the sequence stratigraphic model are
the Late Cretaceous Book Cliffs of Utah, which have previously been described as containing up to 10 sequence boundaries within
the Santonian-Campanian section. These sequence boundaries are defined by channelized erosion surfaces overlain by lowstand and
transgressive, tidally influenced, estuarine valley fill successions that contain stacked tidal point bars, shell banks, coals and some paleosols
EC

(e.g., Davies et al., 2006). However, recent research by Howell et al. (2018) concludes that there is no unequivocal stratigraphic evidence
for periods of relative sea-level fall within the depositional record at this time, and that there are alternative explanations which do
not invoke sea-level change for the formation of these surfaces. Alternative explanations include the accumulation of sediment within
raised mires undergoing differential compaction that result in stacked geometries, sudden influxes of sediment and rapid progradation, or
variations in tidal range on shorelines that have a stronger tidal influence than previously described (Howell et al., 2018; Pattison, 2018).
In this and several other examples, significant and deep incision is shown to occur in response to autogenic sedimentary processes such as
RR

tidal scour, which may be misinterpreted as wave-base erosion during progradation despite there being no evidence for sea-level fall.
In carbonate depositional systems, repetitive and cyclic facies variations in shallow-water carbonate platforms are often attributed to
changes in sea-level or orbital Milankovitch, cyclicity (Weij et al., 2018). It had been suggested by Cozzi et al. (2005) that shallow-water
carbonates completely fill accommodation space during sea-level rise and that sedimentary thickness is a direct measure of sea-level
change. Shallow seismic data from the Great Bahama Bank, where sea-level history, water depth and sediment types are well constrained,
show that accommodation space formed during Holocene sea-level rise is not filled in the way predicted by traditional sequence
CO

stratigraphic interpretations and that there is a poor correlation between sedimentary thickness, facies type and accommodation space (i.e.,
water depth) (Weij et al., 2018). This observation supports studies by Drummond and Wilkinson (1993), Soreghan and Dickinson
(1994) and Wilkinson et al. (1997) which concluded that sedimentary thickness and lithofacies composition are unrelated to water depth
and eustatic sea-level change. Instead, hydrodynamic energy levels along the Great Bahama Bank carbonate platform are shown to be the
primary control on the sedimentary infill of accommodation space and distribution of inner platform facies (Weij et al., 2018). Specifically
mud-rich facies dominate in low-energy areas of the Great Bahama Bank such as the inner platform lagoon, whereas mud-lean facies
UN

dominate in high energy settings where strong currents influence the platform (e.g., Purkis et al., 2017; Weij et al., 2018). Unpredictable
sedimentary thickness of carbonate cycles on the Great Bahama Bank platform and varied hydrodynamic energy levels result in steady and
large off-bank transport of sediment and irregular facies distribution that is independent of water depth within the inner platform (Weij et
al., 2018). This confirms that sedimentary thicknesses of shallow-marine carbonates are unreliable as a direct measure of the amplitude of
orbitally controlled sea-level change and creation of accommodation space (Weij et al., 2018).
The concept of “highstand shedding” in carbonate depositional systems states that during periods of high sea-level, carbonate platform
interiors become flooded. Due to the limited accommodation space, material such as aragonite mud is transported or “shed” into areas
of deeper water to form a sediment wedge of off-platform mixed or allochthonous sediment (Schlager et al., 1994). However, some
carbonate systems, in particular the distally steepened carbonate ramp of the Australian North West Shelf, can produce large amounts of
Sea-Level Change in Geological Time 11

aragonite mud that form lowstand wedges during periods of low sea-level with a volume of sediment that rivals typical highstand systems
(Dix et al., 2005; Reuning et al., 2018).
These examples demonstrate that energy distribution, shelf/delta width and morphology, icehouse vs. greenhouse conditions, eustatic
sea-level change, relative sea-level change (incorporating tectonic subsidence), sediment supply and transport routes are all important
factors on the control of depositional sequences and care must be taken when interpreting sedimentary sequences to account for multiple
variables.

F
Reconstructing Sea-Level Change
Reconstructions of relative sea-level change are based on empirical data that estimate sea-level height during a particular period or range

OO
of geological time. The reconstructed sea-level history is produced by measuring the bathymetry of the sea-level indicator with respect to
a modern datum such as global mean sea-level (Kemp et al., 2015). Sea-level indicators are physical, chemical or biological proxies that
can be shown to occur at a particular bathymetry. In ancient sediments bathymetry must be established by direct measurement of modern
analogues and sea-level indicators dated through correlation with marine oxygen or strontium isotope or magnetic signatures, radiometric
methods (such as potassium‑argon dating of bentonite marker beds) and biostratigraphy (Kemp et al., 2015). The increasing recognition
of the non-uniqueness of sedimentary strata is placing added importance on determining localized relative sea-level curves (e.g., Morley

PR
et al., 2011; Gold et al., 2018) for specific geographic regions, basins or tectonic regimes, rather than relying on global eustatic sea-level
curves to explain the distribution of sediment within a basin. For example, in tectonically active Southeast Asia tectonic events occur
over time intervals below the resolution of global third order sequences so that these sequences cannot be reliably applied. However, if
contemporaneous sea-level events or trends are identified in several different basins, these are likely to be regionally or globally significant
and may record a eustatic signal (Simmons et al., 2007). Some tectonically stable areas such as passive margins or regions with simple,
homogeneous geomorphologies such as broad shallow platforms or carbonate ramps are more susceptible to eustatic sea-level change (e.g.,
Sharland et al., 2001; Di Lucia et al., 2017; Gold et al., 2018) than regions with more complex geomorphologies.

D
Walther’s Law states that conformable vertical changes in facies reflect transitions between adjacent depositional environments.
These transitions in depositional environment may occur through changes in sea-level, changing rates in sediment supply or variation in
environmental conditions. In some instances, lithological features may provide direct evidence for sea-level fluctuations. The most reliable
TE
facies changes that can be used to determine sea-level change are sequences that mark the transition from marine to nearshore to fluvial
depositional environments. The distinction between marine and terrestrial sediments can be determined through the absence or presence of
definitive marine taxa, such as nannofossils, dinoflagellates and marine foraminifera. Terrestrially derived sediments may contain evidence
for land plants, including those that produce miospores, freshwater plants (e.g., charophyte-bearing green algae) and invertebrates, and
land-dwelling vertebrates, for example. Care must be taken to examine all available evidence to determine the environment of deposition, as
EC

miospores produced by land plants are often wind-blown into marine settings; even some dinosaur specimens are known from marine strata
having been washed out to sea. An up-section transition from marine to terrestrial strata may indicate a period of sea-level fall (regression),
or high rates of sediment supply causing the shoreline to prograde seaward. Conversely, an up-section transition from terrestrial to marine
strata may indicate sea-level rise (transgression).
RR

Stable Isotopes
Stable isotope geochemistry, particularly oxygen isotope curves provide an independent proxy of sea-level history (Rovere et al., 2016;
Burgess et al., 2008). Data for the creation of oxygen isotope curves are collected from marine carbonates, predominantly biogenic
calcite from the shells of organisms including molluscs, corals and foraminifera, and silica. Oxygen has three naturally occurring isotopes,
16O, 17O, and 18O, which occur on Earth in the approximate proportions 99.757%, 0.038%, and 0.205%, respectively. The isotope 18O

has a greater atomic mass and is therefore heavier than 16O. Traditionally, changes in isotope composition are denoted relative to the
CO

heavier isotope. With oxygen isotopes this is described as δ18O, which represents the difference (per mil, ‰) in the ratio of 18O/16O of a
sample compared to the Vienna Standard Mean Ocean Water (V-SMOW) or international carbonate isotope (V-PDB) standards. During
non-glacial periods, ocean water is well mixed with both heavy (18O) and light (16O) oxygen isotopes, and has a V-SMOW value of − 1.2‰
δ18O. The oxygen isotope 16O is more easily evaporated from seawater as it is lighter than 18O; this is known as a fractionation process.
Consequently, fresh (or meteoric) water is enriched with 16O and depleted in 18O, resulting in more negative values of δ18O (Fig. 8). An
example of this phenomenon is observed when comparing the oxygen isotope composition of tropical waters which have typical δ18O
UN

values up to − 1‰ V-SMOW and seawater from polar regions which has more negative δ18O values up to − 3‰ V-SMOW.
The trend of δ18O values for marine carbonates through geological time displays positive and negative oxygen isotope excursions. It
is inferred that positive oxygen isotope excursions indicate glacial periods and low global sea-level when much 16O is locked up in ice
sheets and the oceans more enriched in 18O, resulting in less negative δ18O values (Fig. 8). Conversely, a negative oxygen isotope excursion
indicates warmer periods where fresh water is introduced back into the oceans from the melting ice sheets, causing global sea-level rise and
diluting seawater with 16O, resulting in highly negative δ18O values (Fig. 8).
Consequently, eustatic sea-level curves that reflect oxygen-isotope trends (Fig. 9) are more reliable than those that rely solely on coastal
onlap. Important negative oxygen isotope excursions associated with eustatic sea-level rise occur in the Late Cretaceous, Eocene, and
Neogene (Fig. 9). Determining global sea-level patterns using oxygen isotope records works well for the Cenozoic and Mesozoic, however,
well preserved Paleozoic marine fossils, reef cements and cherts give relatively low δ18O values, even in non-glacial times.
12 Sea-Level Change in Geological Time

F
OO
PR
D
TE
Fig. 8 Fractionation process of oxygen isotopes within the Earth’s oceans. Isotopically light 16O is preferentially evaporated from seawater and is
EC

locked up as ice during periods of low sea-level resulting in positive oxygen isotope excursions where oceans are enriched in the heavier isotope,
18O. During greenhouse periods, ice sheets melt and dilute seawater with 16O resulting in negative oxygen isotope excursions and periods of high

sea-level.
RR

The value of δ18O within a foraminiferal test records the oxygen isotope ratio of the seawater from which it was precipitated and the
temperature of calcification (Pearson, 2012). The values of oxygen isotope ratios from calcite and the interpreted isotopic value of the
seawater in which it was precipitated (e.g., − 1.2‰ SMOW estimated for Cretaceous seawater Shackleton and Kennett, 1975) can be
used to calculate the temperature of ancient seawater using the famous paleo-temperature equation of Epstein et al. (1953), which has
subsequently been modified by several authors. Oxygen isotope ratios from the tests of deep sea benthic foraminifera have been used to
CO

identify global climate variations over the past 100 million years, while ratios from the tests of planktonic foraminifera indicate the surface
temperatures of ancient oceans (Pearson, 2012).
Emiliani (1954) used oxygen isotope ratios from the tests of planktonic foraminifera to establish that different isotopic ratios within
the tests reflected the different depth habitats of the foraminifera (Fig. 10). This research has been used to constrain interpretations
of paleobathymetry from fossil assemblages in stratigraphic sections that contain ancient foraminifera analogous to present-day forms.
Using oxygen isotope ratios to calculate paleo-temperature from the tests of planktonic foraminifera collected from Pleistocene to Recent
UN

sediments, Emiliani (1955) later identified a series of warm and cool periods which formed the first observations of glacial and inter-glacial
cycles throughout the Quaternary. Emiliani also examined oxygen isotope ratios from the tests of benthic foraminifera from sediments of
different ages to establish that the temperature of the seafloor has also varied with time, providing evidence for global cooling throughout
the Cenozoic.
More recent applications of oxygen isotope ratios from the tests of foraminifera as paleo-thermometers have seen research into climate
change through deep geological time. Deep water benthic foraminiferal oxygen isotope ratios have been used to highlight how ocean
temperature and ice volume has varied considerably since the Early Cretaceous (Pearson, 2012). Significant ice sheets were largely
absent during the Early Eocene (Cramer et al., 2011), following the Paleocene-Eocene Thermal Maximum (PETM), when seafloor
temperatures were approximately 14°C, over 10°C higher than today (which are currently 2–3°C), and global sea-level was approximately
180 m higher than today (Snedden and Liu, 2010). Global temperatures may have been even higher during the “Middle” Cretaceous
(Pearson, 2012), and continuing research into oxygen isotopes of foraminifera deposited during the PETM, Eocene-Oligocene transition
Sea-Level Change in Geological Time 13

F
OO
PR
D
TE
EC
RR
CO
UN
14 Sea-Level Change in Geological Time

and Middle Miocene Climatic Optimum is providing high resolution data on variations of past climate and global sea-level during these
times.

Biological Sea-Level Indicators


By far the best way to determine past changes in sea-level is through paleontological data. Present day faunal assemblages, and the mode of
life and morphology of organisms, can be used to determine the depositional environment of ancient rocks that contain comparable fossils.

F
Spores, pollen, algae, diatoms, testate amoebae, nannofossils, radiolaria and foraminifera are among the organisms that have been used
successfully to reconstruct past sea-level change (e.g., Barlow et al., 2014; Saher et al., 2015; Gold et al., 2017a, 2018). Foraminifera
are not only useful in oxygen isotope analyses, they can also be used as a proxy for cool- or warm-water conditions and as direct depth

OO
indicators to constrain paleobathymetry in marine sediments. Interpretations of past climatic conditions from micropaleontological evidence
may be achieved through observations of the coiling directions of certain planktonic foraminifera, such as Globorotalia truncatulinoides
(Fig. 11). It is interpreted that the dextrally (to the right) coiling forms of G. truncatulinoides may prefer warmer water conditions, and
sinistrally (to the left) coiled forms prefer colder waters (Bé and Tolderlund, 1971; Fig. 11). Based on observations of the coiling direction
of G. truncatulinoides throughout a succession of Pleistocene strata, interpretations of the stratigraphic position of past interglacial or
glacial episodes, and respective periods of high and low sea-level, can be made. However, the presence of cooler water oceanic currents

PR
(Bé and Tolderlund, 1971) or provincialism arising from genetic variation between different populations (Ujiié et al., 2010) may provide
alternative hypotheses for the presence of sinistrally coiled morphotypes in sediments otherwise interpreted to have been deposited during
warmer climates. In determining paleobathymetry larger benthic foraminifera are particularly useful for shallow-water carbonates, whereas
small calcareous benthic and planktonic foraminifera are of more use in deeper water strata.
Many studies use the modern and ancient depth distribution of foraminiferal taxa as potential sea-level or water depth indicators (e.g.,
Hallock and Glenn, 1986; Leckie and Olson, 2003; Hohenegger, 2005; Gold et al., 2017a, 2018; Goeting et al., 2018). The morphology
of many different types of foraminifera can be used to determine water depth, hydrodynamic energy and substrate type. Many larger

D
benthic foraminifera, such as nummulitids, lepidocyclinids, and orthophragminids, occur in shallow, warm water carbonate platforms.
These foraminifera house photosynthetic algal symbionts which indicate deposition in oligotrophic conditions and at water depths within
the photic zone (variable, but up to approximately 200 m), the zone at which sunlight can still penetrate the water column. Most miliolid
TE
foraminifera are also oligotrophic and have thick microgranular walls which protect them from harmful ultraviolet radiation in shallow,
clear water. Some smaller benthic foraminifera such as Amphistegina house photosynthetic symbionts adapted to blue light that can
penetrate greater water depths of up to 150 m, although healthy populations of Amphistegina thrive in water depths between 20 and 30 m
(Goeting et al., 2018). Amphistegina is also commonly associated with sandy sediments in warm, tropical environments with their robust
tests able to withstand high energy. There is a general trend for an increase in the diameter of flat, discoidal larger benthic foraminifera with
EC

water depth (e.g., Hallock and Glenn, 1986). This is similar to the strategies of many species of platy coral (Fig. 12). These foraminifera
and coral grow large to increase the surface area available to capture sunlight with increasing water depth. Morphogroup types of small
calcareous benthic foraminifera can also be used as bathymetric indicators. Some examples of key indicator genera include Melonis,
Chilostomella, and Uvigerina which are most abundant in outer shelf to upper bathyal settings (the replacement of U. mediterranea with U.
peregrina occurs at about at about 1500 m), Globobulimina is most abundant in upper bathyal settings, while the absence of Globobulimina
and Chilostomella, and presence of Hoeglundia and Oridorsalis may signify a middle bathyal setting (Phipps et al., 2012). Other smaller
RR

benthic foraminifera can be found in bathyal environments especially where they are agglutinating and not made of calcite, and some
miliolids occur in deep water where nutrients are also scarce.
Planktonic foraminifera are usually only found in the inner to outer neritic (continental shelf) environment and rarely within lower
bathyal environments due to the “Carbonate Compensation Depth,” a depth below which carbonate is dissolved; typically this boundary is
found between 4 and 5 km. The majority of planktonic foraminifera spend most of their lives feeding within the photic zone, with highest
concentrations found between 10 and 50 m (Bé and Tolderlund, 1971; Bé, 1977). In addition, many juvenile specimens of planktonic
CO

foraminifera occur within epipelagic, shallow water depths, moving to deeper waters later in their life cycles. As discussed previously,
Emiliani (1954) used oxygen isotope ratios from the tests of planktonic foraminifera collected at different depths within the water column
can be used to calculate the temperature of precipitation of the tests which can then be compared against the measured water column
temperature profile (Pearson, 2012; Fig. 10). The depth stratification of species determined in this way has been validated by many
other studies which collected planktonic foraminifera from plankton nets and sediment traps (e.g., Birch et al., 2013; Meilland et al.,
2019). These studies aid the identification of several ecological groups which can be used as analogues for fossil forms to determine
UN

ancient variations in bathymetry. Generally, planktonic foraminifera which inhabit shallower water depths are small with thin-walled tests
when compared to those that inhabit deeper water as these are larger and have thicker walls, commonly exhibiting keels and/or calcitic
overgrowths. A generalized stratification of present day oceanic water columns can be established based on planktonic foraminiferal

Fig. 9 Oxygen isotope records of Grossman (2012) between the Late Cretaceous and Neogene showing comparable trends in negative δ18O
excursions and increases in global sea-level of the Snedden and Liu (2010) long-term eustatic sea-level trend. Chronostratigraphy modified from
Timescale Creator (version 6.1.2, http://www.tscreator.org) and based on the Geological Time Scale 2012.
Sea-Level Change in Geological Time 15

F
OO
PR
Fig. 10 (Left) Graph displaying percentage of planktonic foraminifera expected with increasing water depth. (Right) Depth habitats of modern

D
foraminifera. Modified from Birch, H., Coxall, H.K., Pearson, P.N., Kroon, D. and O'Regan, M., (2013). Planktonic foraminifera stable isotopes
and water column structure: Disentangling ecological signals. Marine Micropaleontology 101, 127–145, with images planktonic foraminifera from
Ovechkina, M.N., Bylinskaya, M.E. and Uken, R. (2010). Planktonic foraminiferal assemblage in surface sediments from the Thukela Shelf, South
Africa. African Invertebrates, 51(2), 231–255.
TE
EC
RR
CO
UN

Fig. 11 Coiling directions of Globorotalia truncatulinoides. Sinsitrally coiled forms indicate cool water conditions (glacial periods) whereas
dextrally coiled forms indicate warm water conditions (interglacial periods).
16 Sea-Level Change in Geological Time

F
OO
PR
Fig. 12 Environmental preferences and changing morphology of foraminifera and coral on a typical carbonate rimmed platform. Modified from
Hallock, P. and Glenn, E.C. (1986). Larger foraminifera: A tool for paleoenvironmental analysis of Cenozoic carbonate depositional facies.
Palaios, 1, 55–64; Tucker, M.E. and Wright, V.P. (1990). Carbonate sedimentology. John Wiley & Sons.

D
morphology (Bé, 1977). The upper 50 m of the water column is dominated by “shallow water” forms comprising small, simple, thin-walled
and globular (globigerine) morphologies such as species of Orbulina, Globigerina, and Globigerinoides (Fig. 13). Between 50 and 100 m,
TE
“intermediate water” forms comprise more robust, thick-walled, carinate (keeled) and planoconvex morphologies such as species of
Globorotalia, Sphaeroidinella, and Neogloboquadrina (Fig. 13). In water depths in excess of 100 m, “deep water” forms comprise adult
stages with an abundance of carinate morphologies. An assemblage dominated by carinate foraminifera may indicate several hundreds
of meters water depth. Deep planoconvex morphologies of planktonic foraminifera have been suggested to inhabit deeper waters in later
stages of their life cycle.
EC

Occasionally, planktonic foraminifera can be washed into much shallower settings. Investigations into the composition of foraminiferal
assemblages in the shallow-water Pedro Bank, offshore Jamaica, show that that washed-in planktonic foraminifera can contribute up to
20% of the assemblage (Fig. 10). These environments should be interpreted based on the greater number of benthic foraminifera over
planktonics within the fossil assemblage and also by examining other fossil indicators such as changing ichnofacies and a large number
of miospores, most abundant in shallow water settings and with increasing numbers of dinoflagellate cysts further away from shore as
identified from palynological analyses.
RR

Other key indicators for water depths include trace fossils. Trace fossils, or ichnofossils, often represent the burrows, resting or feeding
traces of organisms as they move on or through sediment (Fig. 13). Assemblages of particular types of ichnofossils are called ichnofacies
and often signify a particular water depth, substrate or energy level within a given depositional environment (Fig. 13). The Nereites
ichnofacies contains ichnofossils that predominantly occur on top of the substrate, found as bed-parallel fossils, and often indicate low
energy, deep-water depositional settings (Fig. 13). In contrast, the Skolithos ichnofacies contains ichnofossils that penetrate the substrate,
CO

found perpendicular or oblique to bedding as fossils, and often indicate high energy depositional settings. These high energy settings may
be found in shallow water areas where wave action is high or in deep water areas where turbidity currents provide the necessary high energy
(Fig. 13).

Carbonate Sediments
Carbonates are particularly susceptible to climate change as they are deposited in sensitive and finely balanced biological ecosystems
UN

vulnerable to changes in sea-level, weather, nutrient and light levels, dissolved CO2 and other environmental factors. In carbonate factory
settings, sea-level rise may result in carbonate platform drowning. There are many examples of drowned carbonate platforms documented in
the rock record worldwide (e.g., Brandano et al., 2016; Sulli and Interbartolo, 2016; Gold et al., 2017b). Carbonate platform drowning
occurs because the rate of platform top production and accumulation is exceeded by the rate of relative sea-level rise. This can occur
either because the rate of relative sea-level rise has increased, or because the rate of carbonate production has decreased. However, rates
of carbonate production and accumulation normally exceed rates of long-term sea-level rise and subsidence leading to the “drowning
paradox.” The drowning paradox occurs where carbonate platforms are drowned even though under optimum conditions platform growth
should keep pace or exceed the rate of relative sea-level rise (Table 2). However, the drowning paradox can be explained when the rate of
relative sea-level rise is lower than the maximum carbonate growth potential, and there is no evidence for significant external forcing such
as tectonic activity or glacio-eustatic changes, if carbonate accumulation rates at the initial water depth are lower than the rate of relative
Sea-Level Change in Geological Time 17

F
OO
PR
D
TE
Fig. 13 Distribution of trace fossils and foraminifera along a shallow- to deep-water transect across the continental shelf. Modified from Frey,
EC

R.W., Pemberton, S.G. and Saunders, T.D. (1990). Ichnofacies and bathymetry: A passive relationship. Journal of Paleontology, 64(1), 155–158.

sea-level rise (Kim et al., 2012). This can be induced by slow response times of carbonate-producing marine benthic organisms to
re-colonize the platform top after flooding causing negative feedback in the rates of carbonate accumulation which decreases exponentially
RR

with increasing water depth (Kim et al., 2012). However, the rate of glacioeustatic sea-level rise can indeed exceed carbonate production,
for example during the “Flandrian Transgression” (Donovan and Jones, 1979).
An emerging area of research is the use of rare earth elements (REEs) to identify drowning events in carbonate sediments. The
profile of REE distribution within a stratigraphic section can be used to determine changes in bathymetry as some REEs are associated
with shallow-water depositional environments, whereas others indicate more deep-water settings. For example, low cerium (Ce) and high
yttrium (Y) signatures within a stratigraphic section indicate a deep marine depositional setting. In contrast, a broad samarium (Sm),
CO

europium (Eu), gadolinium (Gd), terbium (Tb), and dysprosium (Dy) signature is associated with terrestrial run-off indicating a proximal
depositional setting close to shore.
Coral reefs have suffered five episodes of severe, global-scale biodiversity loss and termination of reef growth, termed “reef crises”
throughout geologic time (Scheibner and Speijer, 2008; Kiessling and Simpson, 2011; Pandolfi et al., 2011). These reef crises are
defined by global declines in actively produced reefal carbonate volume per unit time and occur during periods of increased sea surface
temperature and pCO2 levels, and declining pH resulting in ocean acidification contemporaneous with rising sea-level (Kiessling and
UN

Simpson, 2011). Coral reefs are particularly sensitive to increasing temperatures as the symbiotic relationship between scleractinian corals
and pigment-giving zooxanthellae breaks down when temperatures are anomalously high, resulting in “coral bleaching” (Pandolfi et al.,
2011). The upper temperature threshold for coral bleaching ranges from 24°C to 34°C (Scheibner and Speijer, 2008), and throughout
much of the Phanerozoic global temperatures were > 7°C higher than today (Pandolfi et al., 2011). Although no direct evidence for coral
bleaching is observed in the fossil record as pigments in corals are not preserved, estimates of global sea surface temperatures during reef
crises suggest that coral bleaching may have taken place and the response of carbonate platforms to these crises are documented (e.g.,
Scheibner and Speijer, 2008; Kiessling and Simpson, 2011).
The five major post-Cambrian metazoan reef crises are identified as follows: (1) Late Devonian, after the Frasnian-Famennian stage
boundary, (2) Early Triassic, after the end Permian mass extinction, (3) very Early Jurassic, after the Triassic-Jurassic boundary, (4) Early
Jurassic, after the Pliensbachian-Toarcian stage boundary and (5) Early Eocene, after the Paleocene-Eocene boundary (Kiessling and
18 Sea-Level Change in Geological Time

Simpson, 2011). The four most recent biodiversity crises in reef ecosystems do appear to coincide with episodes of rapid global warming,
rising sea-level and ocean acidification. These were preceded by elevated extinction rates of corals and sponges (Kiessling and Simpson,
2011; Pandolfi et al., 2011). Evidence suggests that global warming, and the corresponding sea-level rise, was probably the dominant
trigger for the Triassic–Jurassic and the Paleocene-Eocene reef crises (Kiessling and Simpson, 2011). The Early Paleogene experienced
the most pronounced long-term warming trend of the Cenozoic which includes the Paleocene–Eocene Thermal Maximum (PETM), a
brief but significant interval of global warming coincident with a rapid, global negative carbon isotope excursion, which provides the
best evidence for ancient ocean acidification (Scheibner and Speijer, 2008; Kiessling and Simpson, 2011; Robinson, 2011). During the

F
PETM sea-level rose to approximately 180 m higher than the present day (Snedden and Liu, 2010) and tropical Pacific Ocean sea surface
temperatures may have increased by 3.5–5°C, likely exceeding 30°C (Robinson, 2011), well above the threshold for coral bleaching.
Throughout the PETM a marked faunal shift on continental carbonate platforms has been documented, from coral-algal reefs to reefs

OO
dominated by large benthic foraminifera (Scheibner and Speijer, 2008; Pandolfi et al., 2011). As warming progressed, three stages
of carbonate platform development have been identified (Scheibner and Speijer, 2008). During the Late Paleocene, coralgal platforms
consisting of carbonate rock formed by an intergrowth of frame-building corals and algae were dominant at low to middle paleolatitudes
(Stage 1). From the Late Paleocene into Early Eocene, coralgal platforms dominated at middle paleolatitudes as reefs moved towards
the poles where cooler sea surface temperatures were more favorable for reef growth as warming at low latitudes pushed sea-surface
temperatures beyond the maximum temperature range of corals. During this time, platforms dominated by larger benthic foraminifera

PR
such as Miscellanea, Ranikothalia, and Assilina developed at low paleolatitudes at the expense of corals (Stage 2). During the Early
Eocene, platforms dominated by larger benthic foraminifera such as Alveolina, Orbitolites and Nummulites dominated at low to middle
paleolatitudes (Stage 3). The onset of the latter platforms dominated by larger benthic foraminifera correlates with the PETM (Scheibner
and Speijer, 2008). These stages are observed on almost all carbonate platforms attached to continental shelves, however they are not
recorded on thick isolated carbonate platforms that formed away from the influence of continental margins, such as Pacific flat-topped sea
mounts known as “guyots” (Robinson, 2011). The causes for the change from coral-dominated platforms to larger foraminifera-dominated
platforms is linked to the effects of the PETM, resulting in short-term warming, rising sea-level, eutrophic conditions on continental shelves

D
and ocean acidification impeding the growth of aragonitic corals (Scheibner and Speijer, 2008). While corals may have been vulnerable to
symbiont loss at the elevated temperatures associated with the PETM, communities of calcitic larger benthic foraminifera flourished as they
may have been capable of surviving extreme temperatures for a sustained period of time (Robinson, 2011). It is interpreted that the stages
TE
of carbonate platform development observed throughout the PETM may be repeated as the recent global warming and ocean acidification
trend affects today’s coral reefs (Scheibner and Speijer, 2008). A succession displaying a faunal turnover from coral reef dominated to
larger benthic foraminifera dominated may indicate an increase in global paleotemperatures and sea-level rise.

Summary of Determining Sea-Level in the Geological Record


EC

The aforementioned examples show that sea-level rise is intrinsically linked to global temperature rise. In addition, interpretations of the
distribution of sedimentary sequences or stacking patterns alone is not enough to determine past changes in sea-level as the deposition of
these sequences is shown to be controlled by a number of complex factors. Instead, demonstrable sea-level change within a succession
must be identified by several lines of evidence including changes in the bathymetric preferences and ecological tolerances of marine fossil
assemblages, and negative oxygen isotope excursions acting as a proxy for global temperature rise. An idealized succession that contains
RR

evidence for a change in bathymetry, in this example rising sea-level, may include a basal package of strata that contains coarse grained
material such as conglomerates, unidirectional sedimentary structures formed by fluvial processes and a fossil assemblage rich in miospores
and freshwater or brackish organisms signifying a proximal depositional setting. This package may record positive oxygen isotope ratios
and transition to overlying marine-influenced strata. Initially these strata may comprise inner shelf sands or platform carbonates and contain
shallow-water taxa such as larger benthic foraminifera or coral reefs. As the succession continues to deepen upwards, carbonate platforms
drown, forming a drowning succession and drowning unconformity in the process, the oxygen isotope signal becomes increasingly
CO

negative, grain size decreases and more open marine organisms such as pelagic dinoflagellates, nannofossils and planktonic foraminifera
or benthic smaller benthic or agglutinated foraminifera increasingly dominate the fossil assemblage.

Discussion
UN

Research into the response of ancient marine ecosystems to past sea-level rise may help to understand and anticipate the effects of present
day sea-level rise on existing ecosystems. Since the late 19th century, global mean sea-level has been rising. Approximately 1.8 mm/year
of global mean sea-level rise occurred during the latter half of the 20th century, increasing to approximately 3.1 mm/year by the end of
the century (Steffen et al., 2010). Most current research agrees that the rise in global sea-level is driven by thermal expansion of seawater
caused by ocean warming and the melting of ice sheets through anthropogenic climate change. Thermal expansion of ocean water alone
is interpreted to have contributed to 0.4 mm/year of global mean sea-level rise, increasing to 1.5 mm/year during the latter half of the 20th
century (Steffen et al., 2010). Measurements of atmospheric O2 and CO2, which increase as these gases are released from oceans as they
warm, can be used as a thermometer to determine the temperature of the Earth’s oceans (Resplandy et al., 2018). These measurements
suggest that ocean warming, and therefore global sea-level rise, is occurring faster than previously estimated (Resplandy et al., 2018). It
Sea-Level Change in Geological Time 19

is estimated that melting of ice sheets on Antarctica alone may contribute to approximately 60 m of global sea-level rise. Reducing
uncertainty in predicting the future of coral reefs requires a greater understanding of past responses to rapid climate and sea-level change
(Pandolfi et al., 2011). The response of ancient coral reefs during the PETM may serve as a useful analogue for the effects that recent
increases in global mean sea temperatures may have on present day coral reef sites (Scheibner and Speijer, 2008).
Increasing rates of sea-level rise pose a significant hazard to many communities and ecosystems, particularly through more frequent
flooding during storms and high tides (Kemp et al., 2015). In order to mitigate against this hazard, continuing research into ancient
sea-level change is providing new data to help resolve uncertainty over the magnitude and spatial variability of this change aiding prediction

F
of the effects of future sea-level rise at regional and global scales (Kemp et al., 2015). Current models that forecast global mean sea-level
rise do not accurately reflect the expected spatial variability of local sea-level change which may range from sea-level fall to a rise that
outpaces the rate of global mean sea-level change due to a range of physical processes (Kemp et al., 2015). Increasing use of numerical

OO
forward models on ancient analogues will assist in determining the response of depositional settings to past sea-level rise and constrain
which physical processes are likely to be the cause of future hazards.
Advances in ancient sea-level research will likely continue to come from numerical forward modeling and a better understanding of the
non-uniqueness of sedimentary sequences. Recent critical examination of the current state of the sequence stratigraphic model suggests that
new developments to improve stratigraphic understanding and prediction are likely to come from a combination of outcrop, geostatistical
or stochastic approaches, or numerical approaches to the modeling of depositional systems. Numerical stratigraphic forward modeling

PR
experiments enable the refinement of traditional sequence stratigraphic concepts through quantitative assessment of the role of relative
sea-level change and sediment distribution in continental margins (Harris et al., 2018).

References

Barlow, N.L., Long, A.J., Saher, M.H., Gehrels, W.R., Garnett, M.H., Scaife, R.G., 2014. Salt-marsh reconstructions of relative sea-level change in the North

D
Atlantic during the last 2000 years. Quaternary Science Reviews 99, 1–16.
Bé, A.W., 1977. An ecological, zoogeographic and taxonomic review of recent planktonic foraminifera. In: Oceanic micropaleontology. Academic Press, pp.
1–100.
Bé, A.W.H., Tolderlund, D.S., 1971. Distribution and ecology of living planktonic foraminifera in surface waters of the Atlantic and Indian oceans. In:
TE
Micropaleontology of oceans. Cambridge University Press, New York, pp. 105–149.
Birch, H., Coxall, H.K., Pearson, P.N., Kroon, D., O'Regan, M., 2013. Planktonic foraminifera stable isotopes and water column structure: Disentangling
ecological signals. Marine Micropaleontology 101, 127–145.
Brandano, M., Corda, L., Tomassetti, L., Tagliavento, M., 2016. Frequency analysis across the drowning of a lower Jurassic carbonate platform: The Calcare
Massiccio formation (Apennines, Italy). Marine and Petroleum Geology 78, 606–620.
Burgess, P.M., 2006. The signal and the noise: Forward modeling of allocyclic and autocyclic processes influencing peritidal carbonate stacking patterns.
EC

Journal of Sedimentary Research 76 (7), 962–977.


Burgess, P.M., Hovius, N., 1998. Rates of delta progradation during highstands: Consequences for timing of deposition in deep-marine systems. Journal of
the Geological Society of London 155, 217–222.
Burgess, P.M., Wright, V.P., 2003. Numerical forward modeling of carbonate platform dynamics: An evaluation of complexity and completeness in carbonate
strata. Journal of Sedimentary Research 73 (5), 637–652.
Burgess, P.M., Steel, R.J., Granjeon, D., Hampson, G.J., Dalrymple, R.W., 2008. Stratigraphic forward modeling of basin-margin clinoform systems:
Implications for controls on topset and shelf width and timing of formation of shelf-edge deltas. In: Recent advances in models of siliciclastic
RR

shallow-marine stratigraphy. vol. 90, SEPM (Society for Sedimentary Geology), pp. 35–45.
Carvajal, C.R., Steel, R.J., 2006. Thick turbidite successions from supply-dominated shelves during sea-level highstand. Geology 34, 665–668.
Castelltort, S., Honegger, L., Adatte, T., Clark, J.D., 2017. Detecting eustatic and tectonic signals with carbon isotopes in deep-marine strata, Eocene Ainsa
Basin, Spanish Pyrenees. Geology 45, 15–18.
Catuneanu, O., 2006. Principles of sequence stratigraphy. Elsevier, Amsterdam, 375 pp..
Catuneanu, O., et al., 2009. Towards the standardization of sequence stratigraphy. Earth-Science Reviews 92, 1–33. https://doi.org/10.1016/j.earscirev.2008.
10.003.
CO

Covault, J.A., Graham, S.A., 2010. Submarine fans at all sea-level stands: Tectono-morphologic and climatic controls on terrigenous sediment delivery to the
deep sea. Geology 38, 939–942.
Cozzi, A., Hinnov, L.A., Hardie, L.A., 2005. Orbitally forced Lofer cycles in the Dachstein limestone of the Julian Alps (northeastern Italy). Geology 33 (10),
789–792.
Cramer, B.S., Miller, K.G., Barrett, P.J., Wright, J.D., 2011, 23pp., doi:10.1029/2011JC007255. Late Cretaceous–Neogene trends in deep ocean temperature
and continental ice volume: Reconciling records of benthic foraminiferal geochemistry (δ18O and Mg/Ca) with sea level history. Journal of Geophysical
Research, Oceans 116 (C12023), .
Davies, R., Howell, J., Boyd, J., Flint, S.S., Diessel, C., 2006. High-resolution sequence-stratigraphic correlation between shallow-marine and terrestrial
UN

strata: Examples from the Sunnyside Member of the Cretaceous Blackhawk Formation, Book Cliffs, eastern Utah. The American Association of Petroleum
Geologists Bulletin 90, 1121–1140.
Di Lucia, M., Sayago, J., Frijia, G., Cotti, A., Sitta, A., Mutti, M., 2017. Facies and seismic analysis of the Late Carboniferous–Early Permian Finnmark
carbonate platform (southern Norwegian Barents Sea): An assessment of the carbonate factories and depositional geometries. Marine and Petroleum
Geology 79, 372–393.
Dix, G.R., James, N.P., Kyser, T.K., Bone, Y., Collins, L.B., 2005. Genesis and dispersal of carbonate mud relative to late quaternary sea-level change along
a distally-steepened carbonate ramp (Northwestern Shelf, Western Australia). Journal of Sedimentary Research 75, 665–678.
Donovan, D.T., Jones, E.J.W., 1979. Causes of world-wide changes in sea level. Journal of the Geological Society 136 (2), 187–192.
Drummond, C.N., Wilkinson, B.H., 1993. Aperiodic accumulation of cyclic peritidal carbonate. Geology 21, 1023–1026.
Emiliani, C., 1954. Depth habitats of some pelagic foraminifera as indicated by oxygen isotope ratios. American Journal of Science 252, 149–158.
Emiliani, C., 1955. Pleistocene temperatures. Journal of Geology 63, 538–578.
20 Sea-Level Change in Geological Time

Epstein, S., Buchsbaum, R., Lowen-Stam, H.A., Urey, H.C., 1953. Revised carbonate-water isotopic temperature scale. Geological Society of America
Bulletin 64, 1315–1325.
Goeting, S., Briguglio, A., Eder, W., Hohenegger, J., Roslim, A., Kocsis, L., 2018. Depth distribution of modern larger benthic foraminifera offshore Brunei
Darussalam. Micropaleontology 64, 299–316, no. 4, text-figures 1–3, tables 1–4, plates 1–2.
Gold, D.P., White, L.T., Gunawan, I., BouDagher-Fadel, M.K., 2017. Relative sea-level change in western New Guinea recorded by regional biostratigraphic
data. Marine and Petroleum Geology 86, 1133–1158.
Gold, D.P., Burgess, P.M., BouDagher-Fadel, M.K., 2017. Carbonate drowning successions of the Bird’s head, Indonesia. Facies 63 (4), 25.
Gold, D.P., Fenton, J.P., Casas-Gallego, M., Novak, V., Pérez-Rodríguez, I., Cetean, C., Price, R., Nembhard, N., Thompson, H., 2018. The biostratigraphic
record of cretaceous to Paleogene tectono-eustatic relative sea-level change in Jamaica. Journal of South American Earth Sciences 86, 140–161.

F
Grossman, E.L., 2012. Oxygen isotope stratigraphy. In: Gradstein, F.M., Ogg, J.G., Schmitz, M., Ogg, G. (Eds.), The geologic time scale 2012. Elsevier.
Hallock, P., Glenn, E.C., 1986. Larger foraminifera: A tool for paleoenvironmental analysis of Cenozoic carbonate depositional facies. PALAIOS 1, 55–64.
Haq, B.U., Al-Qahtani, A.M., 2005. Phanerozoic cycles of sea-level change on the Arabian platform. GeoArabia 10 (2), 127–160.

OO
Hardenbol, J., Thierry, J., Farley, M.B., Jacquin, T., De Graciansky, P.-C., Vail, P.R., 1998. Mesozoic and Cenozoic sequence chronostratigraphic framework
of European basins. In: In: de Graciansky, P.-C., et al. (Eds.), Mesozoic and Cenozoic sequence stratigraphy of European basins. pp. 109–127, SEPM
Spec. Publ, no. 60, SEPM (Society for Sedimentary Geology), Tulsa, Oklahoma .
Harris, A.D., Covault, J.A., Madof, A.S., Sun, T., Sylvester, Z., Granjeon, D., 2016. Three-dimensional numerical modeling of eustatic control on
continental-margin sand distribution. Journal of Sedimentary Research 86, 1434–1443.
Harris, A.D., Baumgardner, S.E., Sun, T., Granjeon, D., 2018. A poor relationship between sea level and deep-water sand delivery. Sedimentary Geology
370, 42–51.
Hohenegger, J., 2005. Estimation of environmental paleogradient values based on presence/absence data: A case study using benthic foraminifera for

PR
paleodepth estimation. Palaeogeography, Palaeoclimatology, Palaeoecology 217, 115–130.
Howell, J.A., Eide, C.H., Hartley, A., 2018. No evidence for significant sea level fall in the cretaceous strata of the book cliffs of eastern Utah. In: ACE 2018
Annual Convention & Exhibition.
Kemp, A.C., Dutton, A., Raymo, M.E., 2015. Paleo constraints on future sea-level rise. Current Climate Change Reports 1, 205–215.
Kiessling, W., Simpson, C., 2011. On the potential for ocean acidification to be a general cause of ancient reef crises. Global Change Biology 17 (1), 56–67.
Kim, W., Fouke, B.W., Petter, A.L., Quinn, T.M., Kerans, C., Taylor, F., 2012. Sea-level rise, depth-dependent carbonate sedimentation and the paradox of
drowned platforms. Sedimentology 59 (6), 1677–1694.
Kominz, M.A., Browning, J.V., Miller, K.G., Sugarman, P.J., Mizintseva, S., Scotese, C.R., 2008. Late cretaceous to Miocene Sea-level estimates from the

D
New Jersey and Delaware coastal plain coreholes: An error analysis. Basin Research 20 (2), 211–226.
Le Heron, D.P., Armstrong, H.A., Wilson, C., Howard, J.P., Gindre, L., 2010. Glaciation and deglaciation of the Libyan desert: The late Ordovician record.
Sedimentary Geology 223 (1–2), 100–125.
Leckie, R.M., Olson, H.C., 2003, 5-19, SEPM Spec. Publ, no. 75, SEPM (Society for Sedimentary Geology), Tulsa, Oklahoma. Foraminifera as proxies for
TE
sea-level change on siliciclastic margins. In: Olson H. C. and Leckie R. M., (Eds.), Micropaleontologic proxies for sea-level change and stratigraphic
discontinuities.
Meilland, J., Siccha, M., Weinkauf, M.F., Jonkers, L., Morard, R., Baranowski, U., Baumeister, A., Bertlich, J., Brummer, G.J., Debray, P., Fritz-Endres, T.,
2019. Highly replicated sampling reveals no diurnal vertical migration but stable species-specific vertical habitats in planktonic foraminifera. Journal of
Plankton Research 41 (2), 127–141.
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., Pekar, S.F., 2005.
EC

The Phanerozoic record of global sea-level change. Science 310, 1293–1298.


Morley, R.J., Swiecicki, T., Pham, D.T.T., . A sequence stratigraphic framework for the Sunda region, based on integration of biostratigraphic, lithological
and seismic data from Nam Con Son basin, Vietnam. In: Proceedings, Indonesian Petroleum Association, 35th Ann. Convention, Jakarta, May 2011.
Pandolfi, J.M., Connolly, S.R., Marshall, D.J., Cohen, A.L., 2011. Projecting coral reef futures under global warming and ocean acidification. Science 333
(6041), 418–422.
Pattison, S.A., 2018. Rethinking the Incised-Valley fill paradigm for Campanian book cliffs strata, Utah–Colorado, USA: Evidence for discrete
Parasequence-scale, Shoreface-Incised Channel fills. Journal of Sedimentary Research 88 (12), 1381–1412.
RR

Pearson, P.N., 2012. Oxygen isotopes in foraminifera: Overview and historical review. The Paleontological Society Papers 18, 1–38.
Phipps, M., Jorissen, F., Pusceddu, A., Bianchelli, S., De Stigter, H., 2012. Live benthic foraminiferal faunas along a bathymetrical transect (282–4987 m) on
the Portuguese margin (NE Atlantic). Journal of Foraminiferal Research 42 (1), 66–81.
Purkis, S., Cavalcante, G., Rohtla, L., Oehlert, A.M., Harris, P.M., Swart, P.K., 2017. Hydrodynamic control of whitings on great Bahama Bank. Geology 45,
939–942.
Resplandy, L., Keeling, R.F., Eddebbar, Y., Brooks, M.K., Wang, R., Bopp, L., Long, M.C., Dunne, J.P., Koeve, W., Oschlies, A., 2018. Quantification of
ocean heat uptake from changes in atmospheric O2 and CO2 composition. Nature 563 (7729), 105.
CO

Reuning, L., Hallenberger, M., Gallagher, S., Belde, J., Back, S., Fulthorpe, C.S., 2018. The quaternary carbonate system of the north west shelf of Australia:
Lowstand wedge and Highstand bulge. In: Geological Society of London Seismic characterisation of carbonate platforms and reservoirs conference
abstracts, 10th–11th October, London.
Robinson, S.A., 2011. Shallow-water carbonate record of the Paleocene-Eocene thermal maximum from a Pacific Ocean guyot. Geology 39 (1), 51–54.
Rovere, A., Stocchi, P., Vacchi, M., 2016. Eustatic and relative sea level changes. Current Climate Change Reports 2 (4), 221–231.
Saher, M.H., Gehrels, W.R., Barlow, N.L., Long, A.J., Haigh, I.D., Blaauw, M., 2015. Sea-level changes in Iceland and the influence of the North Atlantic
oscillation during the last half millennium. Quaternary Science Reviews 108, 23–36.
Scheibner, C., Speijer, R.P., 2008. Late Paleocene–early Eocene Tethyan carbonate platform evolution—A response to long-and short-term paleoclimatic
UN

change. Earth-Science Reviews 90 (3–4), 71–102.


Schlager, W., Reijmer, J.J., Droxler, A., 1994. Highstand shedding of carbonate platforms. Journal of Sedimentary Research 64 (3b), 270–281.
Shackleton, N.J., Kennett, J.P., 1975. Paleotemperature history of the Cenozoic and the initiation of Antarctic glaciation: Oxygen and car-bon isotope analyses
in DSDP sites 277, 279, and 281. Initial Reports of the Deep Sea Drilling Project 29, 743–755.
Sharland, P.R., Archer, R., Casey, D.M., Davies, R.B., Hall, S.H., Heward, A.P., Horbury, A.D., Simmons, M.D., 2001. Arabian plate sequence stratigraphy,
GeoArabia special publication 2, Gulf PetroLink, Bahrain, 371 p..
Simmons, M.D., Sharland, P.R., Casey, D.M., Davies, R.B., Sutcliffe, O.E., 2007. Arabian plate sequence stratigraphy: Potential implications for global
chronostratigraphy. Geoarabia 12 (4), 101–130.
Snedden, J.W., Liu, C., 2010. A compilation of Phanerozoic Sea-level change, coastal onlaps and recommended sequence designations, AAPG Search &
Discovery Article 40594.
Soreghan, G.S., Dickinson, W.R., 1994. Generic types of stratigraphic cycles controlled by eustasy. Geology 22, 759–761.
Steffen, K., Thomas, R.H., Rignot, E., Cogley, J.G., Dyurgerov, M.B., Raper, S.C., Huybrechts, P., Hanna, E., 2010. Cryospheric contributions to sea level
rise and variability. In: Understanding sea-level rise and variability. Wiley-Blackwell, pp. 177–225.
Sea-Level Change in Geological Time 21

Stow, D., 2012. Vanished Ocean: How Tethys reshaped the world. Oxford University Press.
Sulli, A., Interbartolo, F., 2016. Subaerial exposure and drowning processes in a carbonate platform during the Mesozoic Tethyan rifting: The case of the
Jurassic succession of Western Sicily (Central Mediterranean). Sedimentary Geology 331, 63–77.
Sweet, M.L., Blum, M.D., 2016. Connections between fluvial to shallow marine environments and submarine canyons: Implications for sediment transfer to
deep water. Journal of Sedimentary Research 86, 1147–1162.
Thorne, J.A., In: Watkins J. S., Zhiqiang F. and McMillan K. J., (Eds.), Geology and Geophysics of Continental Margins, AAPG Memoir 53,1992, 375-394,
American Association of Petroleum Geologists, Tulsa, Oklahoma. An analysis of the implicit assumptions of the methodology of seismic sequence
stratigraphy: Chapter 21: Sea-level and seismic stratigraphic studies.
Ujiié, Y., de Garidel-Thoron, T., Watanabe, S., Wiebe, P., de Vargas, C., 2010. Coiling dimorphism within a genetic type of the planktonic foraminifer

F
Globorotalia truncatulinoides. Marine Micropaleontology 77 (3–4), 145–153.
Walsh, J.P., Nittrouer, C.A., 2003. Contrasting styles of off-shelf sediment accumulation in New Guinea. Marine Geology 196, 105–125.
Weij, R., Reijmer, J.J., Eberli, G.P. and Swart, P.K., 2018. The limited link between accommodation space, sediment thickness, and inner platform facies

OO
distribution (Holocene-Pleistocene, Bahamas). The Depositional Record, 2018, 1-21, doi: 10.1002/dep2.50
Wilkinson, B.H., Drummond, C.N., Rothman, E.D., Diedrich, N.W., 1997. Stratal order in peritidal carbonate sequences. Journal of Sedimentary Research
67, 1068–1082.
Zhang, J., Steel, R., Olariu, C., 2017. What conditions are required for deltas to reach the shelf edge during rising sea level?. Geology 45 (12), 1107–1110.

Further Reading

PR
Gradstein, F.M., Ogg, J.G., Schmitz, M., Ogg, G.e., 2012. The geologic time scale 2012. Elsevier.
Kominz, M.A., 2001. Sea level variations over geologic time. Encyclopedia of ocean sciences. Academic Press, San Diego, 2605–2613. https://doi.org/10.
1006/rwos.2001.0255.
Krauskopf, K.B., Bird, D.K., 1994. Introduction to geochemistry, 3rd edn. McGraw-Hill, New York.

D
Glossary
TE
Anthropogenic Factors controlled by man-made forces.
Aphelion The farthest distance from the Earth to the Sun during an elliptical orbit.
Bathymetry The relief of the Earth’s surface that is below mean sea-level (submarine).
Carbonate factory Carbonate factories are carbonate depositional settings which have variations in the dominant precipitation
mode (e.g., abiotic, biotically induced or biotically controlled), mineral composition, depth range and growth potential of carbonate
EC

production. There are five major carbonate factories which include: warm and shallow (tropical), cold (temperate-polar), pelagic (open
ocean), microbial and non-marine.
Dinoflagellate A type of cyst producing aquatic algae, commonly found in marine environments as plankton but are also common in
freshwater. Blooms of dinoflagellates may produce bioluminescence or toxic red tides. Dinoflagellate cysts are commonly used within
biostratigraphy.
Eccentricity Eccentricity describes the degree of ellipticallity of the Earth’s orbit around the Sun. Eccentricity produces the lowest
RR

order, longest term, Milankovitch or orbital cycle, resulting in sea-level oscillations occurring over a period of approximately
100,000 years.
Eustatic Eustatic sea-level is measured from a fixed reference point, often the center of the Earth. Eustatic sea-level change occurs
slowly, globally and synchronously often as a result of the expansion and retreat of ice sheets and glaciers at the poles.
Falling stage systems tracts Sedimentary sequences interpreted to be deposited during the falling, regressive, limb of a sea-level
cycle.
CO

Foraminifera Foraminifera are a group of single-celled amoeboid protists that belonging to the Rhizaria supergroup. These
organisms often secrete a shell, or test, formed of calcium carbonate or agglutinated particles and are characterized by pseudopodia
that have a granular texture to the cytoplasm which is connected by small holes, or foramen, in the walls of the test. Foraminifera are
often described as benthic, living on or near the seabed, or planktonic where they are free floating in the water column.
Fractionation A separation process in which a quantity of a material is divided into a smaller quantity, or fraction, whereby the
composition of the fraction is controlled by a gradient.
UN

Highstand shedding Highstand shedding occurs in carbonate depositional systems during periods of high sea-level when platform
interiors become flooded and due to the limited accommodation space material is transported, or “shed,” into areas of deeper water
forming a sediment wedge of off-platform mixed or allochthonous sediment.
Highstand A period of time when sea-level is at its highest, often representing the maximum transgressive inflection of the sea-level
cycle.
Highstand systems tracts Sedimentary sequences interpreted to be deposited during periods of high sea-level.
Hydrological cycle A cycle whereby water is transported around the surface of the Earth as a solid, liquid or gas through natural
processes. An example includes the evaporation of warm liquid seawater which then cools causing water vapor to condense and form
clouds which later precipitate rain.
Isostatic Localized sea-level change occurring in response to tectonic forces such as isostatic loading.
22 Sea-Level Change in Geological Time

Large igneous province A large igneous province (LIP) is a large accumulation of intrusive and extrusive igneous rocks produced
by the emplacement of magma and eruption of lava.
Lowstand A period of time when sea-level is at its lowest, often representing the maximum regressive inflection of the sea-level
cycle.
Lowstand systems tracts Sedimentary sequences interpreted to be deposited during periods of low sea-level.
Maximum flooding surface A stratigraphic surface produced during highstand conditions at the maximum transgressive reflection
of the sea-level curve representing the deepest water environment at the time of deposition.

F
Mid-ocean ridge Mid ocean ridges occur at the site of seafloor spreading and form vast submarine mountain ranges due to the
uplifting of the crust as magma is extruded to the seafloor. The formation of these ridges is often responsible for the displacement of

OO
much seawater causing sea-level rise.
Milankovitch cycles Sea-level cycles that are controlled by factors relating to the Earth’s spin on its axis and orbit around the Sun,
namely eccentricity, obliquity and precession. Defined by Serbian scientist Milutin Milanković in the early 20th century.
Nannofossils Calcareous nannofossils that comprise nanometer scale coccoliths, coccospheres and nannoliths. Coccospheres are
formed by coccolithophores, a type of phytoplankton, and are the major biological component that forms chalk. Nannofossils are one
of the common index fossils used in biostratigraphic dating and are most abundant from the Triassic onwards.
Obliquity Refers to the variation in the degree of tilt of the Earth’s axis. The tilt of the Earth’s axis varies between 22.1° to 24.5° (an

PR
angle of 2.4°) in cycles of approximately 41,000 years.
Oligotrophic Water conditions that comprise low nutrient availability often defined in terms of micrograms of chlorophyll or
phosphate per liter volume of water (0.0–2.6 μg/L chlorophyll, 0.0–12.0 μg/L phosphate).
Panthalassa A large ocean that existed between the Proterozoic and early Mesozoic. It was the precursor to today’s Pacific ocean
and at its peak during the Paleozoic-Mesozoic transition covered 70% of the Earth’s surface.
Perihelion The closest distance from the Earth to the Sun during an elliptical orbit.
Plate tectonics The Earth’s crust is divided into numerous fragments, or plates, which move in relation to each other through a

D
variety of processes including seafloor spreading and subducted slab-pull.
Precession Refers to the ‘wobble’ of the Earth’s axial rotation which occurs due to the gravitational pull of the Sun and the Moon on
the water at Earth’s equatorial bulge of water occurring in cycles of around 26,000 years.
TE
Rare earth elements (REEs) One of a set of seventeen chemically similar metallic elements. Despite their name, these elements are
not particularly rare but tend to occur together in nature and are difficult to separate from one another.
Regression Occurs where there is net seaward movement of the shoreline when the rate of sediment supply exceeds the rate of
sea-level rise or when sea-level is falling.
Relative sea-level Relative sea-level change is measured from a position on the Earth’s crust, such as the continental shelf, whereby
EC

relative sea-level rise may occur in response to local tectonic subsidence.


Sequence boundary A stratigraphic surface produced when subaerial exposure forms an erosive unconformity and related
correlative conformity.
Stratigraphic forward model A fully quantitative, deterministic three-dimensional computer model that simulates the depositional
evolution of stratal sequences and predicts the thickness and spatial distribution of facies, stacking patterns and stratal geometries.
Models are based on a set of predefined parameters and boundary conditions including, but not limited to, sea-level change, sediment
RR

supply, tectonics and environmental or climatic factors often constrained by seismic and/or well data.
Tethys The Tethys Ocean was originally located in the embayment formed by the supercontinent Pangaea. During the break-up of
Pangaea in the early to middle Mesozoic it spread from east to west and was eventually closed in the Neogene as the continents of
Africa and South America moved north towards Europe and North America, respectively. It existed for approximately 250 million
years between the Permian (256 Ma) and Neogene (5.3 Ma).
CO

Thermodynamics Physical laws that govern the interaction of different forms of energy.
Topography The relief of the Earth’s surface that is above mean sea-level.
Transgression Occurs where there is net landward movement of the shoreline when the rate of sediment supply is constant and
sea-level is rising.
Transgressive systems tracts Sedimentary sequences interpreted to be deposited during the rising, transgressive, limb of a sea-level
cycle.
Unconformity A stratigraphic contact between two rock types of different ages, separated by a significant depositional hiatus.
UN

Walther’s Law Named after the German geologist Johannes Walther and states that a vertical change in facies represents a lateral
change in depositional environment. For example, a shallowing upwards sequence may indicate a shift in depositional environment
from deep-water basin, to a proximal, near-shore, shallow water setting.
Wilson cycle Named after the Canadian geologist John Tuzo Wilson and describes the cyclical nature of the formation and break-up
of continents. A continent initially breaks up by a process of rifting which eventually leads to the creation of a new ocean basin with
a mid ocean ridge at its center. Spreading at the mid ocean ridge pushes material on the fringes of basins together leading to collision
and convergence either resulting in subduction or suturing of the crust and may lead to the formation of a new continent.

You might also like

pFad - Phonifier reborn

Pfad - The Proxy pFad of © 2024 Garber Painting. All rights reserved.

Note: This service is not intended for secure transactions such as banking, social media, email, or purchasing. Use at your own risk. We assume no liability whatsoever for broken pages.


Alternative Proxies:

Alternative Proxy

pFad Proxy

pFad v3 Proxy

pFad v4 Proxy