Reprint - Fachbereich Geowissenschaften

advertisement

Int J Earth Sci (Geol Rundsch) (2010) 99 (Suppl 1):S171–S189

DOI 10.1007/s00531-010-0553-y

O R I G I N A L P A P E R

Quaternary oceans and climate change: lessons for the future?

Wolfgang H. Berger • Michael Schulz •

Gerold Wefer

Received: 3 November 2009 / Accepted: 20 April 2010 / Published online: 16 May 2010

Springer-Verlag 2010

Abstract There is much interest in ice-age studies in recent decades, in the context of global warming. The relevant findings are these: large regular changes in climate occurred within the last million years, especially in the northern North

Atlantic. Extreme conditions were similar, suggesting strong negative feedback at the edges of the range of variation. The nature of the periods of climate variation suggests orbital forcing by modulation of internal oscillations involving lagged negative feedback on ice buildup. Transitions from cold to warm were rapid and they were not readily reversed, indicating that ice dynamics underlies abrupt climate change. Accelerated rates of ice decay upon warming correspond to a sea-level rise of one to two m/century.

Millennial-scale abrupt disturbances known as ‘‘Dansgaard-

Oeschger’’ and ‘‘Heinrich’’ Events occur when large ice masses are present in the northern hemisphere. They may be considered experiments on the ocean’s response to massive meltwater input. When using results from ice-age studies to project future developments, one must be aware that the future will be largely outside of experience with regard to the recent geologic past. Also, there are as yet no generally accepted explanations for striking changes in the past, such as the rapid rise of carbon dioxide during deglaciation, demonstrating a profound lack of understanding of climate dynamics. This is, so far, the lesson from ice-age studies: they say much about the deficiencies in our level of understanding, and not so much about what is ahead.

W. H. Berger

Scripps Institution of Oceanography, University of California,

San Diego, CA 92093-0244, USA

M. Schulz G. Wefer (

&

)

MARUM and Fachbereich Geowissenschaften,

Universita¨t Bremen, 28359 Bremen, Germany e-mail: gwefer@marum.de

Keywords Quaternary Climate Ice-age

Global warming Future

Introduction

Global warming has been proceeding for more than a century, and questions regarding its causes and future course engender much discussion both within the community of Earth scientists and in public discussions. To many researchers in this field, the rapid retreat of mountain glaciers across the globe is the most striking manifestation of the warming (Fig.

1

). Glaciers holding information from the early Holocene and beyond are now melting, signifying that the present warming is unprecedented on a time scale of thousands of years (Thompson et al.

2004

).

The rise of sea level that corresponds to the melting of mountain glaciers (Braithwaite et al.

2002

; Raper and

Braithwaite

2005

) is rather modest, compared with the expansion of the water column as the ocean warms

(Levitus et al.

2001

). It is negligible, with regard to the potential rise of sea level from the rapid decay of ice masses in Greenland and Antarctica. This problem of rapid decay (frequently referred to as ‘‘collapse’’) is central to the global warming conundrum, and it is here that geological studies can make crucial contributions. Geophysical processes involving abrupt change (such as ‘‘collapse’’ of a reservoir) are manifestations of threshold effects, which are difficult to treat from first principles. Invariably, we have to turn to experience and statistics to judge the risks; that is, we have to consider history.

The assessment of the risks of sea-level rise involves the response of polar ice caps to warming, and it involves the redistribution of heat and moisture on the planet, and the feedbacks associated with such redistribution.

123

S172 Int J Earth Sci (Geol Rundsch) (2010) 99 (Suppl 1):S171–S189

Fig. 1 Rapid retreat of mountain glaciers in the Canadian Rockies: tongue of the Columbia Ice Field, Banff/Jasper National Park. Fresh morainal deposits cover much of the valley below the remnant ice tongue. They are underlain by ice in places. (Photo W.H.B. 2006.)

Fig. 2 Effects of weakening of the trade winds as seen in sea-level anomalies during the El Nin˜o event of winter 1997. (Adapted from

K }

2002

; data from NOAA). Note the rise of sea levels along eastern boundaries ( orange ), along with a slowing of the turning of the great central gyres, especially in the northern hemisphere ( blue ).

Eastern boundary upwelling is greatly diminished

The experience that can tell us how fast sea level can rise is contained in the marine sediments that accumulated during the last million years. Without doubt, therefore, the main contribution of marine geology to the global warming discussion is to provide constraints on the rise of sea level, along with its various implications for abrupt changes in climate.

In addition, the study of Quaternary sediments has much to offer regarding changes in the ocean’s circulation, the recycling of nutrients, and the response of the ocean’s productivity to cooling and warming. Productivity patterns are part of the marine carbon cycle, which controls the partitioning of carbon dioxide between ocean and atmosphere. Since the present ocean normally holds roughly sixty times more carbon dioxide than the atmosphere

(Sundquist

1985

), modest changes in the holding capacity of the sea have large effects on the content of greenhouse gas in the air.

These various items of vital interest—sea-level change, heat transport, carbon cycle, productivity—are all poorly understood for the present ocean. This is not owing to a shortage of data (data banks are massive and growing) but because of a lack of control on dynamics and sensitivity to disturbance. For example, while the overall patterns of a positive ENSO event are well mapped by satellite observation (Fig.

2

), the basic physical causes are still in doubt

(Wang and Picaut

2004 ). Apparently, trade winds weaken

and fail to pile up warm water in the western equatorial

Pacific, as suggested by Bjerknes ( 1969

) some time ago. In turn, a weakening of the trade winds leads to a weakening of upwelling regimes in eastern tropical regions. A connection between strength of trade winds and intensity of eastern boundary upwelling is also seen when reconstructing productivity patterns for glacial and interglacial periods (Berger et al.

1989 ; Schneider et al.

1996 ). The

123 correspondence is encouraging in that it supports the idea that there are large-scale patterns of ocean behavior that reappear across time scales from millennia to decades

(Fedorov et al.

2006

). Clearly, as suggested at the very outset of paleo-oceanographic studies (Arrhenius

1952 ),

the strength of zonal winds is a crucial item in tracking the behavior of the ocean’s climate system.

In this essay, we attempt to summarize those results of

Quaternary research that are most relevant to the questions arising in the context of global warming. We emphasize marine geologic studies, because that field is the one we are most familiar with, from research projects pursued and conferences convened within the last two decades (Berger

2003

; Berger et al.

1989

,

1991 ; Fischer and Wefer 1999

;

Wefer et al.

1996

,

1998 , 1999

,

2002

,

2003a , b

). We proceed from a defense of the historical approach, to the nature of ice-age fluctuations, to lessons contained therein with respect to global warming.

History matters

Geologically based opinions rest on three claims: (1) What happened can be reconstructed, (2) what happened must be possible, and (3) the response to forcing observed is typical; that is, it repeats for similar events. The first claim rests on the quality of proxy reconstructions. The second acknowledges that models tuned to modern observations may not be able to capture the full range of reconstructed climate variations. The third relies on repetition, that is, on the application of statistics.

The lack of repetition that characterizes the instrumental period of climate research (roughly the last 150–100 years) is a major problem. It means that the range of variability

Int J Earth Sci (Geol Rundsch) (2010) 99 (Suppl 1):S171–S189 S173 measured by instruments may not be relevant to the mode of operation of the system that is being explored. The result is that much uncertainty attaches to the output of relevant models that create scenarios beyond known variability. By adjusting the models to deal with history, their performance can be greatly improved (see examples cited in IPCC

2007

).

The systematic study of climate change began with geological insights, and notably with contemplating the enormous climate changes associated with the ice ages

(Croll

1875

; Penck and Brueckner 1901–1909; Ko¨ppen and

Wegener

1924

; Brooks

1926 ; see summary in Schwarzbach

1974

, p. 6). The concept of ‘‘climate,’’ in meteorology, was originally meant to denote a mean state of weather fluctuations around seasonal and interannual variability. What geologists showed is that the ‘‘mean state’’ is a function of the time span considered.

An important result of this pioneering work spanning many decades is the map showing the great expanse of ice that dominated northern regions some 20,000 years ago

(Fig.

3 ; for more recent reconstructions see Clark et al.

2009

). Enormous climate changes are implied. The very existence of ice sheets built over tens of thousands of years indicates self-stabilization through feedback from the ice caps and from a lowered sea level, while their relatively rapid disappearance points to positive feedback in decay. The repeated buildup and decay emphasizes the great sensitivity of high northern latitudes to climate change (for background see Jackson and Broccoli

2003 ). That this high sensitivity

holds for short time scales, as well, has been amply verified by observation of temperature trends in the last century

(Jones et al.

2001

) and by modeling (Cubash et al.

2001 ).

The key feedbacks that amplify change in the region are the reflectivity of the ground and the moisture in the air, factors that were discussed more than a 100 years ago by the geologist James Croll and the naturalist Alfred Russell

Wallace. Wallace, for example, wrote as follows (

1895 ,

p. 157): … the increased heat of summer could not be in any way stored up, but would be largely prevented from producing any effect, by reflection from the surface of the snow and by the intervention of clouds and fog …

Reflectivity (albedo) is now generally recognized as the dominant feedback factor. The net contributions of clouds and fog, although clearly important, are less obvious and are difficult to quantify (cf. Soden and Held

2006 ).

The systematic reconstruction of ice-age climates based on microfossils in deep-sea sediments goes back to the pioneering work of Schott (

1935 ) of the

Meteor Expedition

(1925–1927) and especially to the scientists working on the long cores collected by the Albatross during the Swedish

Deep Sea Expedition (1947–1948) (Arrhenius

1952

;

Phleger et al.

1953 ; Emiliani

1955 ; Parker

1958

; Olausson

1961

) (Fig.

4 ).

The piston coring device used in the Albatross Expedition, invented by the oceanographer and engineer Bo¨rje

Kullenberg, turned out to be the time machine needed to penetrate deeply into the detailed history of the Quaternary.

The temperature preferences for planktonic foraminifers became central to the deciphering of the climate record, as did the isotopic analysis of foraminifers introduced by

Emiliani (

1955 ) and first applied by him to cores from the

Albatross expedition, as well as to cores from Lamont

Geological Observatory, collected in the 1950s. Like all fossils, foraminifers are not measurement devices but rather ‘‘biased reporters’’ (Wefer and Berger

1991 ). Once

their biases are recognized, however, they give extremely useful information.

The 1970s saw the first serious attempt to quantify global ice-age conditions on the basis of microfossils

(Imbrie and Kipp

1971 ; CLIMAP Project Members

1981 ).

This opened a path to the possibility of testing the skill of emerging climate models against real-world situations whose patterns were not included already within the models. The approach has proved very valuable, especially as the improvement in the reconstructions provided increasingly reliable targets (COHMAP Members

1988

;

GLAMAP 2000; Sarnthein et al.

2003

; MARGO Project members

2009

).

Fundamental concepts of ice-age narratives include the observed range of conditions, and the rates at which they change. Especially in the context of the ongoing climate change, it is clear that the transitions from one state to another provide the focus of interest, and not equilibrium or quasi-equilibrium states.

All such questions imply the use of a detailed time frame, and detailed correlation between different records.

The necessary stratigraphic control relies heavily on the oxygen isotope analysis of foraminifers (Fig.

5

). In addition, the various aspects of past environments are extracted from a host of proxies, including many types of microfossils, and physical and chemical properties of sediments

(see Fischer and Wefer

1999 ).

Pursuing a climate framework relevant to human affairs demands detailed time resolution (tree rings, lake sediments, ice cores, varved sediments in the oxygen minimum zone, for example). Strikingly high rates of change in temperature and atmospheric CO

2 have been discovered in ice cores from Greenland (e.g. Dansgaard et al.

1993

;

Grootes et al.

1993

; Steffensen et al.

2008 ) and from

Antarctica (Petit et al.

1999 ; EPICA Community Members

2004

;

2006 ; Siegenthaler et al.

2005 ). The rise of atmo-

spheric carbon dioxide concentration during terminations is one such much discussed change (Fig.

6 ). The challenges

posed by the discovery of rapid change in climate and in atmospheric composition have not been met, so far, by generally accepted explanations. Global explanations, of

123

S174

Fig. 3 Approximate land and sea ice cover in the northern hemisphere, during glacial time.

Adapted from Flint ( 1971 )

Int J Earth Sci (Geol Rundsch) (2010) 99 (Suppl 1):S171–S189 course, will have to be based on evidence from marine sediments, in addition to that from ice cores. Also, an important role will remain for studies on sediments with intermediate resolution, which are widespread (in contrast to high-resolution records).

Perhaps the most serious drawback, when using insights from Quaternary geology in building projections for the future, is that the future has no analog in the past. Both the concentration of greenhouse gases and their rates of increase are unprecedented in the Quaternary, as far as one can tell from the record. One has to go back more than 20 million years in Earth history to encounter atmospheric

CO

2 reconstructions as great as concentrations expected in the next centuries (Zachos et al.

2008 ). However, the

geography (land–sea distribution, elevation of mountain ranges) differed substantially from the modern pattern.

This renders the interpretation of potential analogs quite difficult.

In what follows, we discuss major elements of the changing climate system, for the last million years, with a view toward their relevance for the ongoing discussions about global warming. We begin with topics in geophysics

(nature of ice-age cycles, patterns of sea-level change, heat transport, abrupt change) and then move to aspects of the carbon cycle, which is intimately linked to a number of other geochemical cycles, through the various controls on productivity of the ocean.

General aspects of the ice-age record

Concerning climate change of the last million years, two iconic images stand out: (1) The maximum extent of ice cover in the northern hemisphere (Flint

1971

; Imbrie and

Imbrie

1979

) and (2) the sequence of oxygen isotopes in foraminifers from deep-sea sediments (Emiliani

1955

;

Shackleton and Opdyke

1973 ; Imbrie et al.

1984 ; Shackleton

et al.

1990

; Berger and Wefer

1992

; Mix et al.

1995 ; Zachos

et al.

2001 ; Lisiecki and Raymo

2005

). The sequences contain periodicities related to changes in astronomically driven insolation patterns, as emphasized by Emiliani

( 1955

), in accord with expectations based on arguments

123

Int J Earth Sci (Geol Rundsch) (2010) 99 (Suppl 1):S171–S189 S175

Fig. 5 The planktonic foraminifer Globigerinoides sacculifer (adult form with a special final chamber). This is one of the best known of all species used in marine Quaternary research, as the main carrier of oxygen isotope information. (SEM image S.I.O.; diameter ca.

400 l m)

Fig. 4 The Swedish research vessel Albatross on its famous coring expedition (1947–1948). Adapted from Pettersson (

1953

) presented by Ko¨ppen and Wegener (

1924 ) and Milankovitch

(

1930

).

As mentioned in the introduction, the extent of northern hemisphere ice sheets during glacial times was enormous

(Fig.

3 ). In fact, these ice masses of approximately

20,000 years ago, now vanished, exceeded the ice masses of present-day Antarctica, whose ice volume represents a potential sea-level change of 70 m. Antarctica remained largely glaciated throughout the ice-age cycles owing to its being centered on the South Pole. The general cooling started 50 million years ago. Three million years ago it reached the appropriate threshold for allowing ice buildup in the northernmost land areas, well away from the pole.

(For background and references on planetary cooling, see

Zachos et al.

2001 , 2008 ). That the northern ice sheets

could grow at all, and maintain themselves over tens of millennia, is a function of positive feedback from albedo, greenhouse effect, and dust in the atmosphere, within the general framework of a cooling planet.

When dark ground is covered by snow and ice, there is great amplification of the change toward cold conditions, and the same is true for the sea, which is dark when open, and reflects light when covered by sea ice. Also, cold air holds little water vapor, which is the most important of the greenhouse gases. The ice cover, then, makes its own climate, which tends to stabilize the ice.

The high sensitivity of northern circumpolar regions to climate change is the basic reason that Milankovitch forcing, when calculated for summer insolation in high northern latitudes, correlates well with major patterns in the ice-age record. When Wladimir Ko¨ppen proposed 65 8 N as the most sensitive latitude to insolation forcing (Milankovitch

1930 ,

p. A138), he had the subarctic feedbacks in mind. Ko¨ppen’s assessment is supported by modern observations. Recent global warming patterns at the surface suggest enhanced sensitivity to current CO

2 forcing at latitudes near 60 N (see

Figure. 3.5 in IPCC

2007 ). The present rapid decrease in

sea ice cover in the Arctic (Stroeve et al.

2007 ; Comiso et al.

2009

) attests to the fact that this subarctic feedback extends farther north, as well.

The high sensitivity of the subarctic climate system provides the amplification to outside forcing that reverberates throughout the global climate. Other highly sensitive elements of the system exist as well, of course. Thus, the albedo of large land areas in the subtropics changes substantially, depending on the balance of precipitation and evaporation. Desert surfaces tend to be highly reflective

(and also provide dust in air that reflects the sunlight), while vegetation cover is dark. Deserts invaded forested areas during glacials (e.g., Dupont

1999 ), contributing to

123

S176

Fig. 6 Antarctic temperature evolution relative to modern values and atmospheric greenhouse-gas concentrations over the past 800,000 years.

Data from ice cores drilled in

Antarctica (after Lu¨thi et al.

2008 ; Loulergue et al.

2008 )

240

220

200

180

300

0

280

260

100 200

Int J Earth Sci (Geol Rundsch) (2010) 99 (Suppl 1):S171–S189

300 400 500 600 700 800

3

0

-3

-6

-9

-12 800

700

600

500

400

300

0 100 200 300 400 500

Age [1000 years before present]

600 700 800 ground albedo and dust load. Regarding the change from plant cover to no cover in desert regions, it is thought that it can be very rapid, due to positive feedback from the trapping of moisture by the vegetation or the lack thereof

(Claussen

2009 ).

In addition to increased albedo and dust load during glacial periods, there is the diminished greenhouse effect from a decrease in carbon dioxide and water vapor in the atmosphere. Together, these three types of feedback— albedo, greenhouse, dust—lock the planet into its glacial state, until the mountains of ice become unstable and decay, in response to favorable orbital conditions, giving way to the interglacial regime.

The fact that the CO

2 content of the glacial atmosphere is greatly reduced relative to warm conditions still awaits a succinct explanation. In the meantime, there is no shortage of hypotheses (see Archer et al.

2000 ; Sigman and Boyle

2000

). In any case, the large size of the carbon reservoir of the ocean together with changes in the storage capacity of the sea and the exchange with the atmosphere will eventually provide the answer, as initially suggested by

Broecker (

1982 ). However, it is unlikely that a single cause

will strongly dominate: the marine carbon cycle is highly complex (see Archer et al.

2000

).

The single most important type of marine record offering evidence for rates of change and for the amplitudes of past climate excursions is provided by oxygen isotope analysis of planktonic and benthic foraminifers. In addition, this type of record is the chief tool for correlation and for age assignments in deep-sea sediments, based on signal matching. A number of ‘‘standard’’ isotopic curves are available for the purpose, starting with the SPECMAP

scale of Imbrie et al. ( 1984 ), which is valid for the last

650 kyr. For the time before that, the readjustments to the

Quaternary time scale of Shackleton et al. (

1990 ) have to

be considered.

The isotopic record shown in Fig.

7

is based on the compilation of Zachos et al. (

2001 ), resampled for 1,000-

year intervals (Berger

2009 ). Two distinct periods are

readily apparent within this record: the early Pleistocene, beginning near 1,800 thousand years ago and lasting to about 900,000 years ago, and the late Pleistocene, from

123

Int J Earth Sci (Geol Rundsch) (2010) 99 (Suppl 1):S171–S189 S177

900,000 thousand years ago to the Holocene. Near

900,000 years ago, there is a shift to a regime with higher average index values, from a mean of 3.8 to one of 4.1. The

‘‘heavier’’ isotopes indicate colder conditions and greater ice mass. The change in mean is accompanied by a shift to larger amplitudes and longer waves in the climate fluctuations. The event is referred to as the ‘‘Mid-Pleistocene

Climate Shift’’ (Clark et al.

2006 ). What we call the

‘‘Milankovitch Chron’’ comprises the last third of the

Quaternary. It is characterized by the familiar ca. 100-kyr periods, which are commonly referred to the changing eccentricity of the Earth’s orbit. The chron bearing the 100kyr cycles happens to be virtually congruent with the time span originally studied by Milankovitch when proposing the hypothesis that ice-age fluctuations are linked to the changing summer insolation in high northern latitudes

(Milankovitch

1930

).

The cause for the Mid-Pleistocene climate shift is not known. Since the shift is permanent, it probably has to do with changing the framework conditions on the planet, not with the orbital forcing per se or with feedbacks from circulation and other reversible factors. A general threshold important for ice buildup was crossed (for example, in the uplift of mountains) during a time of cool northern summers (beginning about 950 kyr ago; von Dobeneck and

Schmieder

1999

, p. 618) and in consequence the ice mass grew by roughly 20 percent. After that, the climate system never returned to its former mode of operation.

Fig. 7 Ice-age cycles of the last two million years, shown in the d

18

O values of benthic foraminifers from deep-sea sediments, as compiled by

Zachos et al. (

2001

) and resampled for 1-kyr resolution.

a Time series. Numbers are marine isotope stages ( MIS ).

The means differ between early and late Pleistocene; a shift took place near 900,000 years ago.

b Nature of the Mid-Pleistocene climate shift regarding the periodicity of climate fluctuations, based on Fourier analysis. A cycle near

41,000 years is present in both early and late section, but a

100,000-year cycle strongly dominates in the last third of the

Pleistocene

123

S178 Int J Earth Sci (Geol Rundsch) (2010) 99 (Suppl 1):S171–S189

In general, the fluctuations in the early Pleistocene are dominated by a cycle near 41 kyr, which persists. But after the shift, a 100-kyr cycle becomes dominant (Ruddiman et al.

1986

,

1989 ).

The shift indicates that the apparent size of the sun

(governed by the changing Earth-Sun distance, hence eccentricity) becomes important when the northern ice sheets are large. Presumably, obliquity alone can no longer do the job of melting the ice and that makes the precession effect increasingly important. A host of mechanisms have been put forward, some quite complex, to explain the shift in periodicity between early and late Pleistocene. Basically, it seems to us, longer periods of fluctuation win out over shorter ones because larger ice sheets have more inertia.

The shift appears because the available periods that interact with the oscillating system are few and are fixed. Thus, as the potential for ice buildup grows, the system adjusts its main oscillation to the most compatible of the available periodicities, moving from 40 to * 80 to 100 kyr (Mudelsee and Schulz

1997 ).

The message of the Great Shift in the central Quaternary is that 900,000 years ago the system crossed a threshold and changed its response to forcing, in ways that would have been difficult or impossible to predict from previous experience.

The topic of unexpected events brings us to that aspect of ice-age fluctuations that is at the same time among the most vexing and the most important issues: the rapid decay of large ice fields. The importance of these events has been recognized for some time (Broecker and van Donk

1970 ).

While the timing of these ‘‘terminations’’ is somewhat predictable owing to a link to orbital forcing (Paul and Berger

1997 ; Raymo 1997 ; Lisiecki and Raymo

2007 ),

the mechanisms and rates involved are as yet poorly understood.

The 100-kyr riddle enigma (which includes the termination riddle) has two aspects. The first is that eccentricity alone has very little power within the Milankovitch paradigm. Eccentricity becomes effective solely through the modulation of seasonal contrast through precession. The second is that the smallest compatible multiple for the precessional periods (23.7 and 19.0) is 95 (factors of 4 and

5, respectively), so 95 is the period we should expect to see in the record if precessional forcing is important. The problem was recognized and investigated by Muller and

MacDonald (

1997 ), who insisted that eccentricity cannot

explain a 100-kyr cycle, and proposed that it should be replaced with a different orbital parameter.

Before taking such a step, it would be well to gain a more complete understanding of the ice dynamics involved in making the 100-kyr cycles. Ice dynamics, we think, are central to the question of the nature of the terminations that define the 100-kyr cycle. Much has been written about the

123 last termination, which includes a cold spell at the halfway point, when ice ceased to melt for about a 1,000 years. The event is well recorded in deep-sea sediments (Jansen and

Veum

1990

). The pause in deglaciation is known as the

‘‘Younger Dryas’’ period (YD), after periglacial deposits in

Denmark, which contain remains of the Arctic flower

Dryas octopetala . The onset and end of the Younger Dryas period are clearly defined in the oxygen isotope stratigraphy of ice cores on Greenland (see Fig.

8

).

The most striking and significant aspect of the pattern discovered by studying ice core records from Greenland is the fact that the climate is highly unstable during glacial conditions, but is quite stable during the Holocene (Alley et al.

2003 ). The contrast signifies that the northern ice

masses were themselves a source of instability when large

(Schulz et al.

1999

). We do not know what made them so; perhaps it had to do with ice masses grounded well below sea level. Such masses exist today in West Antarctica, and it has been suggested that they are vulnerable to warming and sea-level rise. The ‘‘Heinrich’’ events (named for

Heinrich’s

1988

documentation of layers of ice-rafted debris in North Atlantic deep-sea sediments) are thought to indicate sporadic invasions of enormous fleets of icebergs into the North Atlantic (Bond et al.

1992 ). The Dansgaard-

Oeschger events are sudden and sustained warm spells on

Greenland, which apparently followed the release of ice and meltwater to the sea. If so, the question arises why such release should lead to warming, since presumably it interferes with deepwater production, which helps drive heat import into the Nordic realm (e.g. Sarnthein et al.

2001

).

The warm period immediately preceding the YD

(Fig.

8

) is the first step of deglaciation; it initiates the transition into the present interglacial, the Holocene.

Explaining the cold spell that followed this initial largescale melting has been a major challenge. One often-

Fig. 8 Nature of the transition between the last glacial and the

Holocene as seen in the Greenland ice record. YD, Younger Dryas cold spell. Note the instability of glacial conditions (Dansgaard-

Oeschger events, Heinrich events), compared with the stability of the

Holocene. Adapted from Ganopolski and Rahmstorf (

2001 )

Int J Earth Sci (Geol Rundsch) (2010) 99 (Suppl 1):S171–S189 S179 invoked scenario is a slowdown of the production of North

Atlantic Deep Water (McManus et al.

2004

), based on the concept that NADW production is a vital link in the heat transport from low latitudes to high ones in the North

Atlantic (Broecker and Denton

1989

). At some point, this slowdown scenario became the basis for warnings regarding the risk of interfering with the northward transport of heat in the North Atlantic as the planet warms (Broecker

1997

).

We propose that the problem is not to explain why there was a cold spell at a time when enormous ice sheets still covered Canada and Scandinavia. Cold weather is the expectation for such conditions. The problem is to explain how the Boelling/Alleroed—the warm period preceding the Younger Dryas—started in the first place, and how it maintained itself for a 1,000 years during a period of rapid sea-level rise and freshwater input, which presumably did impact circulation in the manner postulated by Broecker and Denton (

1989 ) along the lines of the shutdown

hypothesis.

Apparently, contrary to the shutdown concept, melting of ice and concomitant freshwater input set in motion positive feedback mechanisms that favored additional melting. If so, one has to explain what stopped the melting, not what caused the cooling after it stopped. One possibility is that the system ran out of vulnerable ice that could be melted given the comparatively weak orbital forcing available to do the job (Berger and Jansen

1995

). Other

possibilities exist. Thus, Schulz et al. ( 2002

) suggested that the Boelling/Alleroed is in fact nothing but a Dansgaard-

Oeschger event that was instigated by the declining ice volume during the last deglaciation. A corollary of this hypothesis is that Younger Dryas type events should have occurred in earlier deglaciations. While there is some evidence for such occurrence in the penultimate deglaciation (Sarnthein and Tiedemann

1990

; Lototskaya and

Ganssen

1999

; Siddall et al.

2006

), some studies suggest the opposite (Carlson

2008

). This question is unresolved.

The main point in the story concerning the transition from glacial to postglacial conditions is that there is unstable ice and stable ice, and that we cannot readily tell the difference before the unstable ice shows its true nature by rapidly wasting. The uncertainty is relevant in the context of the future response of the West Antarctic ice sheet to global warming (Oppenheimer and Alley

2004 ).

vary somewhat. Fifteen years ago, in a report sponsored by the U.S. National Academy of Sciences, Revelle et al.

( 1990

) concluded (p. 4) the ‘‘one hundred years from now it is likely that sea level will be 0.5 to 1 m higher than it is at present.’’ The data we have from the Quaternary record identify this assessment as being realistic to conservative.

To obtain estimates on possible rates of sea-level rise from the wasting of polar ice, we analyzed the oxygen isotope record shown in Fig.

7 . We used the last million

years as compatible with the present workings of the climate system. Linking the oxygen isotopic record to sealevel change is a difficult undertaking and fraught with quite large error bars (e.g., Matthews

1990 ; Peltier

2002 ).

For calibration, we use the usually considered 120 m of sea-level rise since the last glacial maximum (Shepard

1963

; Curray

1965 ; Fairbanks 1989

; Alley et al.

2005

) and relate it linearly to the observed range in the oxygen isotope record, over the same time span (Berger

2008

) (see

Fig.

7

). The assumption is that the change in the isotope values is largely driven by changes in the volume of continental ice, and that the residual that is not so driven runs parallel to the main driving factor. The error is thought to be modest (less than ten percent; Sima et al.

2006

).

The range of isotopes over the relevant interval may be taken as 1.5 permil (Fig.

7 ). Thus, when compared with a

sea-level rise of 120 m, a change of isotope values of

0.125 permil corresponds to a change in sea level of 10 m.

What is of interest in the present context are historical conditions similar to the Holocene, that is, roughly the top

5 percent of the 1000 isotope values defining the last million years in 1-kyr steps. In the data set at hand, we use a cutoff of \ 3.4 permil to obtain the relevant subset. For this set of interglacial values, we compile the rates of change (in permil per kyr) between neighboring points to get the pattern of abundance distributions (Fig.

9 ).

Rates of change of sea level

There is no question that sea level is rising, a fact that is well appreciated by geologists studying the flooding of

Venice, in a region that is sinking (see e.g. Seibold and

Berger

1996 , p. 152). Estimates regarding the global effect

Fig. 9 Abundance distribution of rates of change of oxygen isotope values in Holocene conditions during the last million years. Fit of histogram ( gray area ) by eye. Data are those of Zachos et al. (

2001 ),

resampled (Fig.

7 )

123

S180 Int J Earth Sci (Geol Rundsch) (2010) 99 (Suppl 1):S171–S189

The rates of change associated with a rise in sea level

(marked positive) have some large values, with (20% of rise values showing a rise of more than 1 m per century. In accordance with Rohling et al. (

2008 ), the present analysis

suggests that a rise of 1 m/century would hold no surprise.

This provides support for assessments calling on at least

* 1 m of sea-level rise by the end of the present century

(Rahmstorf

2007

; Grinsted et al.

2009 ). It is however

somewhat higher than the estimate of Kopp et al. (

2009 )

who, based on an analysis of past sea-level changes, consider a 56–92 cm/century range for sea-level rise as likely.

The asymmetry in the patterns of rising and falling sea level reflects the fact that large ice masses involved have a component of instability, as mentioned above. According to the isotope record, this instability is potentially greatest during those times of deglaciation when one-third of the vulnerable ice mass is left to melt (that is, when sea level is near 40 m).

The cause of instability is fairly obvious: (1) ice masses contain gravitational energy, which turns into heat when ice flows downhill, (2) heat transfer into the interior of an ice sheet, by meltwater from the surface percolating into cracks where it can refreeze, is a one-way street (downward only), and (3) meltwater can soften the ground below the ice or form puddles and lakes, which will decrease friction at the base and favor downhill gliding. The point is, together with isostatic sinking under the ice load (which brings the base and rims of ice sheets into vulnerable positions below sea level), there are several mechanisms available to accumulate instability through time, until a threshold is reached for rapid deglaciation.

Detailed quantitative treatment of the various factors involved is complicated, and a matter of much discussion

(for background, see Denton and Hughes

1981

; Oerlemans and van der Veen

1984

; Van der Veen

1999

; Alley et al.

2005

). Nevertheless, some general patterns of ice growth and decay, within the Milankovitch chron, readily emerge from the abundance distribution of sea-level positions.

The main feature of the histogram describing the distributions (Fig.

10 ) is that 90 percent of the range is

between sea-level positions of 10 and 115 m. Between these border values, the likelihood of positions seems to stay roughly the same; there is no indication of a central tendency. This suggests that positive feedback enhances movement away from the center. Near the edges of this quasi-rectangular distribution, negative feedback starts to exceed positive feedback, creating a threshold condition.

Occasionally, the threshold is exceeded, perhaps aided by stochastic resonance in times of unusually warm summers

(cf. Benzi and Sutera

1985 ). Apparently it is possible, once

a threshold is crossed, for the sea level to move into an extreme position (marked ‘‘x-warm’’ in Fig.

10

), with its own quasi-rectangular distribution. The pattern, if read

123

Fig. 10 Histogram of sea-level positions, based on the oxygen isotope data of Zachos et al. (

2001

), transposed by assuming a typical glacial-postglacial range of 120 m. Note the extreme position of present sea level correctly, suggests that there is much reluctance of the system to move into a sea-level position higher than

20 m, but that, once the present position is reached, the probability of moving up or down is roughly the same.

Obviously, this is not a state of affairs that favors stability of sea level, upon strong warming.

A more welcome message from this pattern is that it seems to be very difficult indeed to move sea level upward beyond a position 10 m above the present (Berger

2008 ).

There remains, however, a risk that if future greenhousegas concentrations stay high for a sustained period of time that the existing ice sheets may become unstable (Hansen et al.

2008 ).

For times of extreme cold (sea level at 120 m and lower) we do not see the type of quasi-rectangular distribution seen in the x-warm range. Apparently, building up additional ice, once the sea level is below 115 m, is quite difficult. The issue presumably is one of suitable space; that is, of areas with cool summers and at an elevation high enough to bear snow and ice.

The record (Figs.

9

,

10

) suggests that the usual discussions about the relative importance of factors involved in the recent modest sea-level rise (summarized in Warrick and Oerlemans

2001 ) will turn out to be irrelevant to the

problem of major sea-level rise. For high rates one must assume that the dynamics of ice sheet collapse and meltwater effects in the sea will dominate the situation. It is clear that studies of the role of expansion of seawater or of the input from melting mountain glaciers can teach us much about the workings of the present climate system and about the physics of glacier response to warming. However, once serious melting starts such knowledge will be of limited use.

The overall message from this analysis of the available deep-sea data is that, seen against the ice-age background, negative feedback dominates the climate system near the

Int J Earth Sci (Geol Rundsch) (2010) 99 (Suppl 1):S171–S189 S181 boundary conditions (starting near 20 m, in the present context). This helps keep the system stable. However, in the peculiar Holocene-type condition (extra warm, 10 to

?

10 m), there is no preference attached to any one sealevel position: the sea level can go up as easily as down, by several meters. Does a rapid sea-level rise cause additional rise? Apparently this is so, during deglaciation. However, the Holocene is inherently stable, within narrow limits. To address the question whether a rise in Holocene-type conditions stimulates further rise, we check how many of the rise events during Holocene-type conditions (Fig.

9 )

were followed by additional rise. The answer is 18 out of

41 pairs, slightly less than a 50/50 expectation. Thus, the answer, on the thousand-year scale, is ‘‘not necessarily.’’

But to answer the question on the century scale, which is relevant to human affairs, we would have to have much more detailed information on the sea-level fluctuations of the past. Encouraging steps have been taken in that direction (Siddall et al.

2003

; Rohling et al.

2008 ; Waelbroeck

et al.

2008

; Kopp et al.

2009 ).

North Atlantic heat piracy

Fig. 11 Surface currents in the Atlantic and associated temperature anomalies (W, warm; C, cold) as drawn by Dietrich et al. (

1975 ).

Note the cross-equatorial flow of warm surface water and the role of

Gulf Stream transport in bringing warm water from the subtropics to high latitudes

It has been suggested that global warming might lead to a decrease in the heat supply to the northern North Atlantic, because of a slowdown or shutdown of North Atlantic

Deep Water (NADW) production (e.g., Broecker

1997 ).

When considering deglaciation, however, we found no evidence in the record to support this idea. There is no question that heat transport to the northern North Atlantic is important, of course. That heat imported from far to the south warms northern Europe is well appreciated for more than a century (Croll

1875

; Wallace

1895

) and has been an integral part of introductory oceanography for many decades (Sverdrup et al.

1942

; Dietrich et al.

1975

). The possibility that climate change impacts such transport of heat definitely is a concern.

The topics of changes in thermohaline circulation and implications for heat transport have long intrigued physical oceanographers (Stommel

1961

,

1980 ) and paleoceanog-

raphers (Duplessy et al.

1980

; Crowley

1992 ; Clark et al.

2002

; McManus et al.

2004

; Gherardi et al.

2005 ). More

recently, these topics have generated much interest in modeling, as computing power has grown (e.g. Marotzke

2000

; Rahmstorf

2001

; Paul and Scha¨fer-Neth

2003 ).

Traditionally, the Gulf Stream is cited as a crucial element in the heat transport system, especially with a view to subarctic regions in the North Atlantic, and this connection is deeply engrained in popular thinking about climate (e.g., in describing the warming of the coast of Norway). This is not entirely wrong (Fig.

11 ), although the nature of the

Gulf Stream as part of a wind-driven gyre implies that most of the water masses involved soon return south, rather than moving northward (Stommel

1966 ). Some portion of those

water masses, however, does move north to replenish the waters used in NADW production.

The basic element of North Atlantic heat piracy is the transport across the equator, by both winds and currents, and the general transport northward of heat from the tropics and subtropics within the North Atlantic. Zonal temperature distributions within the Atlantic (see Sverdrup et al.

1942

, p. 127) readily show the overall pattern, with the

South Atlantic distinctly colder than the North Atlantic at all latitudes (Fig.

12 ).

In recent modeling experiments on heat transport by ocean and atmosphere, the transport of latent heat by winds has been identified as the dominant mechanism changing sea-surface temperature in the North Atlantic, on a decadal time scale (Seager et al.

2000

). These results suggest that winds may be more important than ocean currents in moving heat and thus in producing the enormous temperature anomaly of the northern North Atlantic.

However, there is no question that the heat content of the mixed layer in the sea introduces an element of stability in the positioning of high and low pressure regions, so that winds tend to stay with the paths mapped out by currents.

The feeding of heat to the Iceland Low, through moist winds from the Gulf region, and aided by the Gulf Stream, plays a special role in the dynamics of the present climate.

123

S182

Fig. 12 Zonal averages of surface temperatures in the

Atlantic Ocean. Based on

Table 31 in Sverdrup et al.

(

1942

). Note that the maximum difference to southern latitude zones is at 60–70 N. The northward heat transport features contributions from the agulhas current (Ag), the benguela current (Ben), the south equatorial current (C Eq), and the general motion of warm water and moist warm winds toward the intertropical convergence zone (ITCZ).

Within the North Atlantic, the

Gulf stream (Gulf), the westerly winds and currents in temperate latitudes (West) and the Iceland low are involved in heat transport

Int J Earth Sci (Geol Rundsch) (2010) 99 (Suppl 1):S171–S189

The Iceland Low acts as a turntable moving heat into northwestern Europe. The strength of this transport varies on a decadal scale, producing the ‘‘North Atlantic Oscillation.’’ The mechanisms causing the oscillation are not known, nor have the periods expressed therein received much attention (as evidenced in the symposium edited by

Hurrell et al.

2003

). As concerns the conditions during the last glacial maximum, we can be confident that the Iceland

Low was greatly weakened by the surrounding sea ice, by the expansion of the Greenland High, and the general southward movement of the polar front (Fig.

13 ).

In an important but poorly known review, the meteorologist Flohn (

1985 ) has enumerated many implications of

the great planetary climatologic asymmetries linked to northern heat piracy. In his view, the buildup of northern ice caps should lead to greater symmetry in heat distribution, as the strengthened zonal winds north of the equator push the ITCZ toward the equator. It is difficult to tell, from reconstruction of the conditions of the last glacial maximum (LGM), whether or not there was considerable heat transport across the equator (Otto-Bliesner et al.

2009

).

According to the reconstructions of Scha¨fer-Neth and

Paul ( 2003

), their Fig.

1

), it appears that the temperature anomaly of the North Atlantic vanished beyond 40 N (as expected from the southward advance of the polar front) but that heat transfer from south to north persisted in the tropical regions (aided by strong southeasterly winds off the

Amazon; see Paul and Scha¨fer-Neth

2003 , their Fig.

2 ). On

balance, Flohn’s concept, that northward heat transport should be greatly decreased during glacials by centering the

ITCZ closer to the equator, does not seem to be strongly supported by the available data in the Atlantic (Paul and

Scha¨fer-Neth

2003

, p. 568, their Figs.

12

and

13 ).

Fig. 13 Reconstruction of the glacial Atlantic, according to CLIMAP

1984 (as given in Wefer et al.

1999

). The arrow (here added) shows the path of heat transport suggested by the temperature contours (W, warm; C, cold). The eastern South Atlantic is seen as a heat collector for the western tropical Atlantic, similar to the present situation

If cross-equatorial transport of heat in the Atlantic is not highly sensitive to the presence or absence of northern ice sheets (Fig.

13 ), there is little reason to believe that it

would be highly sensitive to the ongoing global warming.

123

Int J Earth Sci (Geol Rundsch) (2010) 99 (Suppl 1):S171–S189 S183

Nevertheless, the difference in patterns (though small) suggests a slight increase in heat piracy of the North

Atlantic, as warming proceeds. This change can be rationalized by considering that warming is stronger on the northern hemisphere with its large land area, and that this excess warming, on the whole, should strengthen monsoon-type import of heat-bearing winds from the south.

For the northern North Atlantic, projections about heat transport are commonly tied to expectations about the response of NADW production to rapid warming and supply of freshwater. The comparison of the last glacial maximum with Holocene conditions, while interesting, would seem to be of only modest relevance to this problem. To a lesser extent, the same is true for the response of NADW production to sustained and rapid meltwater influx during deglaciation. There is no question that massive freshwater input at high and sustained rates must interfere with deep vertical mixing and hence with thermohaline circulation (Olausson

1965

; Berger et al.

1977

; Broecker and Denton

1989 ; Fairbanks 1989

; Manabe and Stouffer

1995

; Stocker and Marchal

2000

;

Rahmstorf

2001

; Tarasov and Peltier

2005

). The question is at what rate and over what period of time does the system enter the state at which stratification wins out over convection in high latitudes. The answer is not clear: it depends on the behavior of a complex set of nonlinear feedbacks.

When assessing the lessons from Quaternary studies for future heat flow patterns, it is well to recall that until very recently computations based on fairly large data sets for the modern ocean ( massive when compared with geologic data) yielded but tentative results for meridional heat flow in the South Atlantic (Schlitzer

1996

). Much progress has been made since, especially with regard to incorporating geologic information (e.g., Paul and Scha¨fer-Neth

2003 ),

but error bars are still substantial.

Carbon cycle and productivity

How the ocean will react to the increasing partial pressure of CO

2 in the atmosphere is clearly a major issue. It is linked to the temperature increase because both an increase in temperature of upper waters, and a decrease in vertical mixing fostered by the warming, will, on the whole, decrease the proportion of carbon dioxide that enters the sea. The chemistry itself—reaction of the carbon dioxide with the water and with the carbonate ion—also leads to a lessening of uptake (as recognized half a century ago by

Revelle and Suess

1957

). Calculations are made difficult not so much from a lack of understanding the basic elements of the chemistry involved, but from the intricacies of vertical circulation and gas exchange at the surface (which determine physical carbon fluxes) and the large number of communicating reservoirs in the biological carbon cycle.

The necessary data to assess the rates of export of carbon from surface waters to deeper layers, in response to climate change, are still rather sparse, although much has been

learned since Suess ( 1980

) reviewed the pertinent problems

(Wefer

1989 ; Wefer and Fischer 1993

; Fischer and Wefer

1996

; Usbeck et al.

2003 ).

The problem of projecting the response of the ocean to

CO

2 values well above present is greatly complicated by the fact that there is no past analog for the present carbon dioxide content of the atmosphere, as far as that can be determined, for the last 800,000 years (see Petit et al.

1999

;

EPICA,

2004 ). Since the atmosphere is a relatively small

reservoir and is intimately linked to a large one—the ocean—the uncertainties concerning the expected atmospheric level, for different conditions postulated for the ocean, are rather large. The rise from lower carbon dioxide values in glacial periods (around 190 ppm) to higher values in interglacials (around 290 ppm) is remarkably rapid and follows closely on the temperature changes seen in the ice

(Fig.

14 ). These, in turn, are highly correlated with the

Fig. 14 Deuterium and carbon dioxide values from terminations in the Vostok ice core (Petit et al.

1999

) mapped on Termination I to show the similarities and differences between these events. During warming the rise of CO

2 follows a regular pattern. (Exceptional value circled ) (From Berger

2003 )

123

S184 Int J Earth Sci (Geol Rundsch) (2010) 99 (Suppl 1):S171–S189 oxygen isotope record in deep-sea sediments (Berger et al.

1996

). It has proved very difficult indeed, so far, to account for the change of 100 ppm in the carbon dioxide content of the atmosphere, across the ice-age cycles. In addition, one must account for another 40 ppm of change that is hidden in the growth of forests and associated soil carbon pools, which change with the climate cycles.

It is unlikely that the mix of mechanisms responsible for increasing carbon dioxide values in atmosphere, at the time of deglaciation, will soon be reconstructed with great confidence (see, e.g., Sigman and Boyle

2000

; Monnin et al.

2001

). However, the message from the record is clear: The system reacts to warming and to a rise in sea level with an increase in atmospheric carbon dioxide.

Another lesson is that equilibrium thinking is unlikely to contribute very much to an appreciation of what the future might bring. The action is with rapid change and its consequences.

Concerning the response of the ocean’s productivity to global warming, the same uncertainties arise as for the carbon cycle, as the two topics are intimately linked.

Changes in the wind field—a weakening of the trade winds, for example—will greatly influence the patterns of upwelling. Based on observational data, Bakun (

1990 )

proposed that global warming due to greenhouse gases should increase upwelling globally. A reconstruction of upwelling intensity off northwest Africa (McGregor et al.

2007

) supports this inference.

Changes in upwelling will affect the several interacting nutrient cycles. In turn, the changes engendered will result in changes in composition of the plankton produced. It is important to realize that large phytoplankton (diatoms) support a food web completely different from that supported by small phytoplankton (such as bacterioplankton or coccolithophores). High-energy consumers (marine mammals and birds) rely on short food chains to provide for an efficient transfer of energy from photosynthesis to consumption (Ryther

1969 ). As a result, we see marine

mammals and birds on seasonal migrations to feeding grounds where diatoms are abundant (Berger

2007

).

Geologists studying the record of productivity have come up against a paradox, well illustrated in the upwelling sediments off Angola and Namibia: Here, glacial sediments contain more marine organic matter than interglacial ones, but show a deficiency of silica relative to warm periods.

Based on the conflicting results of Diester-Haass (

1985 );

more siliceous fossils during warm periods) and Oberha¨nsli

(

1991

); increased glacial upwelling based on foraminifers), it appears unavoidable to postulate a change in nutrient content in upwelled waters, between glacial and interglacial periods, in agreement with Hay and Brock (

1992 ). A

general pattern of silicate deficiency during glacial times may be indicated, for the Namibia upwelling region

123

(Berger and Wefer

2002

; ‘‘Walvis Paradox’’). The Namibia upwelling system may be considered a client of the

Southern Ocean, which ultimately delivers the nutrients for upwelling (Sarmiento et al.

2004

). Thus, changes in the

(sub-)Antarctic silica distribution system seem to be involved.

However, similar strange patterns (increased production, less silica) were discovered in the northern Pacific (Berger and Lange

1998

) and in the eastern equatorial Pacific

(Ganeshram and Pedersen

1998

). This complicates things and suggests that a general shortage of silica arose, because of increased silica deposition in early glacial upwelling systems. A down-draw of atmospheric CO

2 during increased silica deposition in upwelling regions would be expected from the linking of silica and carbon cycles in the sea

(Berger

1991 ; Archer and Maier-Reimer 1994 ; Ragueneau

et al.

2000 ).

From these considerations, it appears that global changes in nutrient content of the thermocline are just as important as changes in the rates of upwelling, as factors controlling ocean productivity.

As concerns the likely effect of warming on ocean productivity, it is apparently closely tied to the response of the trade wind system. In the record, in any case, strong trade winds during glacial times produce increased upwelling both in coastal regions and along the equator.

On the whole, trade winds are a response to the planetary temperature gradient. As this gradient weakens in the northern hemisphere (because of the differential warming of high northern latitudes) we should expect a corresponding weakening of the northern trade winds. Predators that depend on high food density (such as marine mammals and sea birds) would then suffer from a general decrease in the focusing of ocean productivity. Fisheries depending on herring-like fish that strain upwelling waters for food also will suffer, beyond the damage done by overfishing.

We realize that this type of conclusion is embarrassingly general, given the rich background information on production history (or rather proxy history) in many regions of the ocean. However, when proxies disagree, as they do in the Walvis Paradox complex, our knowledge of nutrient cycles and ecosystem dynamics is called upon to resolve the conundrum. This knowledge is as yet deficient, as illustrated by the lively discussions regarding the role of iron in controlling the productivity of large parts of the ocean (Martin and Fitzwater

1988

, and many since). Some have maintained that, rather than iron deficiencies, it is rapid grazing that prevents phytoplankton from depleting the surface waters. Obviously, these are not minor differences in opinion. If one can have a discussion on which is more important, nutrient control or grazing, the system is not well understood.

Int J Earth Sci (Geol Rundsch) (2010) 99 (Suppl 1):S171–S189 S185

Conclusions

We set out to assess which of the many discoveries about climate change in the Late Quaternary (roughly the last million years) would help us most in projecting the response of the climate system to the ongoing overall warming. The (preliminary and tentative) results of this survey are as follows:

• From Quaternary studies, we learned a great many things about the workings of the system, over a large range of climatic conditions, but we did not find an analog to the present or future situation.

• The most instructive periods are those dominated by rapid change, especially from cold to warm, since our own time is a transition of that kind.

• Rapid transitions are driven by strong positive feedback, which is delivered by various effects of warming, including changing albedo patterns, greenhouse-gas increase, sea-level rise, freshwater input, and changes of the wind field and corollaries.

• The last 10,000 years were unusually stable as far as climate change. The critical factor for introducing instability seems to be the presence of ice that is ready to collapse. Historically, when the sea level was in the position that it occupies at present, it could readily move up or down by a meter or two within a century, for several centuries.

• The steps of the last deglaciation have a message that once a threshold is reached, sea level rises rapidly for some time (as long as vulnerable ice is present). The role of the ocean in providing feedback is much discussed and quite poorly understood.

• The North Atlantic will likely continue to draw heat from south of the equator, and store much of it in the

Caribbean realm. In northern subtropical regions, upwelling productivity will likely suffer because of the decrease in zonal winds that drive upwelling.

• The East Antarctic ice sheet is fairly stable. Thus, the winds around the Antarctic are expected to persist, and also the high seasonal productivity that accompanies the deep mixing provided by such winds.

• Warming engenders CO

2 increase (based on ice core results), and there is a strong possibility that diatom productivity (that is, the marine silica cycle) is vitally involved in controlling CO

2 changes. However, neither the carbon cycle nor the silica cycle is well understood.

Perhaps the most obvious conclusion is that a strong link between marine geologic research and other activities (ice coring, climate modeling) can greatly enhance our understanding of the climate system. But we should not expect immediate answers from the past concerning the questions for the future. We should use the experience gathered from climate history and climate modeling to make commonsense projections of what is ahead and find ways to deal with the risks emerging, even in the face of great uncertainty.

Sea-level rise is a phenomenon subject to runaway, based on the history of deglaciation. There is no question that marine-based ice is more vulnerable to destruction than ice safely anchored on solid dry ground. As sea level slowly rises, more and more of the ice becomes marinebased. At some point, further rise, by removing support from marine-based ice, can trigger substantial surging, and also open access for seawater to the base of previously grounded ice. The task, obviously, is to avoid any threshold inviting the start of runaway dynamics.

Acknowledgments We thank William W. Hay and Christian Dullo, who read the article and made useful suggestions for improvement.

References

Alley RB, Marotzke J, Nordhaus WD, Overpeck JT, Peteet DM,

Pielke RA, Pierrehumbert RT, Rhines PB, Stocker TF, Talley

LD, Wallace JM (2003) Abrupt climate change. Science

299:2005–2010

Alley RB, Clark PU, Huybrechts P, Joughin I (2005) Ice-sheet and sea-level changes. Science 310:456–460

Archer D, Maier-Reimer E (1994) Effect of deep-sea sedimentary calcite preservation on atmospheric CO

2

367:26–263 concentration. Nature

Archer DA, Winguth A, Lea D, Mahowald N (2000) What caused the glacial/interglacial atmospheric pCO

2

12:159–189 cycles? Rev Geophys

Arrhenius GOS (1952) Sediment cores from the east Pacific. Rep

Swed Deep Sea Exped 1947–1948, 5:1–228

Bakun A (1990) Global climate change and intensification of Coastal

Ocean Upwelling. Science 247:198–201

Benzi R, Sutera A (1985) The mechanism of stochastic resonance in climate theory. In: Ghil M, Benzi R, Parisi G (eds) Turbulence and predictability in geophysical fluid dynamics and climate dynamics. North-Holland, Amsterdam, pp 403–423

Berger WH (1991) Produktivita¨t des Ozeans aus geologischer Sicht:

Denkmodelle und Beispiele. Zt Deutsch Geol Gesellschaft

142:149–178

Berger WH (2003) Climate future in a warming world: lessons from the ice ages. In: Wernli RL, Kennel CF (conveners), Celebrating the Past … teaming toward the future, Oceans 2003, CD ROM

Proceedings. Holland Enterprises, Escondido, CA, pp 404–408

Berger WH (2007) Cenozoic cooling, Antarctic nutrient pump, and the evolution of whales. In: Warnke D, Filippelli GM, Flores J-A

(eds) Paleoceanography and Paleoclimatology of the Southern

Ocean: a synthesis of three decades of scientific ocean drilling.

Deep-Sea Res II 54:2399–2421

Berger WH (2008) Sea level in the late quaternary: patterns of variation and implications. Int J Earth Sci 97:1143–1150

Berger WH (2009) Ocean—reflections on a century of exploration.

UC Press, Berkeley, p 519

Berger WH, Jansen E (1995) Younger Dryas episode: ice collapse and superfjord heat pump. In: Troelstra SR, van Hinte JE, Ganssen

GM (eds) The younger dryas. North Holland, Amsterdam, pp 61–105

123

S186 Int J Earth Sci (Geol Rundsch) (2010) 99 (Suppl 1):S171–S189

Berger WH, Lange CB (1998) Silica depletion in the thermocline of the glacial North Pacific: corollaries and implications. Deep-Sea

Res II 45:1885–1904

Berger WH, Wefer G (1992) Klimageschichte aus Tiefseesedimenten—Neues vom Ontong Java Plateau (Westpazifik). Die

Naturwissenschaften 79:541–550

Berger WH, Wefer G (2002) On the reconstruction of upwelling history: Namibia upwelling in context. Mar Geol 180:3–28

Berger WH, Johnson RF, Killingley JS (1977) ‘‘Unmixing’’ of the deep-sea record and the deglacial meltwater spike. Nature

269:661–663

Berger WH, Smetacek VS, Wefer G (eds) (1989) Productivity of the ocean: present and past. Wiley, New York, pp 341–354

Berger WH, Kroenke LW, Mayer LA, Shipboard Scientific Party

(1991) Ontong Java Plateau, Leg 130: synopsis of major drilling results. Proc ODP Init Reports 130:497–537

Berger WH, Bickert T, Yasuda MK, Wefer G (1996) Reconstruction of atmospheric CO

2 from the deep-sea record of Ontong Java

Plateau: the Milankovitch chron. Geol Rundschau 85:466–495

Bjerknes J (1969) Atmospheric teleconnections from the equatorial

Pacific. Month Weather Rev 97:163–172

Bond G, Heinrich H, Huon S, Broecker WS, Labeyrie L, Andrews J,

McManus J, Clasen S, Tedesco K, Jantschik R, Simet C (1992)

Evidence for massive discharges of icebergs into the glacial

Northern Atlantic. Nature 360:245–249

Braithwaite RJ, Zhang Y, Raper SCB (2002) Temperature sensitivity of the mass balance of mountain glaciers and icecaps as a climatological characteristic. Z Gletscherkunde und Glazialgeol

38:35–61

Broecker WS (1982) Ocean chemistry during glacial time. Geochim

Cosmochim Acta 46:1689–1705

Broecker WS (1997) Thermohaline circulation, the Achilles heel of our climate system: will man-made CO balance? Science 278:1582–1588

2 upset the current

Broecker WS, Denton GH (1989) The role of ocean-atmosphere reorganizations in glacial cycles. Geochim Cosmochim Acta

53:2465–2501

Broecker WS, van Donk J (1970) Insolation changes, ice volumes and the O18 record in deep-sea cores. Rev Geophys Space Phys

8:169–198

Brooks CEP (1926) Climate through the ages: a study of the climate factors and their variations. E Benn Limited, London, p 439

Carlson AE (2008) Why there was not a Younger Dryas-like event during the Penultimate Deglaciation. Quat Sci Rev 27:882–887

Clark PU, Pisias NG, Stocker TF, Weaver AJ (2002) The role of the thermohaline circulation in abrupt climate change. Nature

415:863–869

Clark PU, Archer D, Pollard D, Blum JD, Rial JA, Brovkin V, Mix

AC, Pisias NG, Roy M (2006) The middle Pleistocene transition: characteristics, mechanisms, and implications for long-term changes in atmospheric pCO

2

. Quat Sci Rev 25:3150–3184

Clark PU, Dyke AS, Shakun JD, Carlson AE, Clark J, Wohlfarth B,

Mitrovica JX, Hostetler SW, McCabe AM (2009) The last glacial maximum. Science 325:710–714

Claussen M (2009) Late Quaternary vegetation-climate feedbacks.

Climate of the Past 5:203–216

CLIMAP Project Members (1981) Seasonal reconstructions of the

Earth’s surface at the Last Glacial Maximum. Geol Soc Am Map and Chart Series MC-36, p 18

COHMAP Members (1988) Climate changes of the last 18, 000 years: observations and model simulations. Science 241:1043–1052

Comiso JC, Parkinson CL, Gersten R, Stock L (2009) Accelerated decline in the Arctic sea ice cover. Geophys Res Lett 35:L01703.

doi: 01710.01029/02007GL031972

Croll J (1875) Climate and time. Appleton & Co, New York

Crowley TJ (1992) North Atlantic deep water cools the Southern

Hemisphere. Paleoceanography 7:489–497

Cubash U, Meehl GA et al. (2001) Projections of future climate change. In: Houghton JT, Ding Y and others (eds) Climate

Change 2001. The scientific basis. Intergovernmental panel on climate change, Cambridge University Press, Cambridge, pp

525–582. (See Figure 9.10.)

Curray JR (1965) Late Quaternary history, continental shelves of the

United States. In: Wright HE, Frey DG (eds) The quaternary of the United States. Princeton University Press, Princeton, pp 723–

735

Dansgaard W, Johnsen SJ, Clausen HB, Dahl-Jensen D, Gundestrup

NS, Hammer CU, Hvidberg CS, Steffensen JP, Sveinbjo¨rnsdottir

AE, Jouzel J, Bond G (1993) Evidence for general instability of past climate from a 250-kyr ice-core record. Nature 364:218–220

Denton GH, Hughes TJ (eds) (1981) The last great ice sheets. Wiley,

New York, p 484 (Maps)

Diester-Haass L (1985) Late Quaternary sedimentation on the eastern

Walvis Ridge, SE Atlantic (HPC 532 and four piston cores). Mar

Geol 65:145–189

Dietrich G, Kalle K, Krauss W, Siedler G (1975) Allgemeine

Meereskunde, Eine Einfu¨hrung in die Ozeanographie. Gebr

Borntra¨ger, Stuttgart, p 593

Duplessy J-C, Moyes J, Pujol C (1980) Deep water formation in the

North Atlantic Ocean during the last ice age. Nature 286:479–

482

Dupont LM (1999) Pollen and Spores in Marine Sediments from the

East Atlantic—a View from the Ocean into the African

Continent. In: Fischer G, Wefer G (eds) Use of proxies in

Plaeoceanography: examples from the South Atlantic. Springer,

Berlin, pp 523–546

Emiliani C (1955) Pleistocene temperatures. J Geol 63:538–578

EPICA Community Members (2004) Eight glacial cycles from an

Antarctic ice core. Nature 429:623–628

EPICA Community Members (2006) One-to-one coupling of glacial climate variability in Greenland and Antarctica. Nature 444:195–

198

Fairbanks RG (1989) A 17, 000 year glacio-eustatic sea level record:

Influence of glacial melting rates on the Younger Dryas event and deep-ocean circulation. Paleoceanography 342:637–642

Fedorov AV, Dekens PS, McCarthy M, Ravelo AC, deMenocal PB,

Barreiro M, Pacanowski RC, Philander SG (2006) The Pliocene

Paradox (mechanisms for a permanent El Nin˜o). Science

312:1485–1489

Fischer G, Wefer G (1996) Long-term observation of particle fluxes in the eastern Atlantic: Seasonality, changes of flux with depth and comparison with the sediment record. In: Wefer G et al (eds)

The South Atlantic: present and past circulation. Springer,

Berlin, pp 324–344

Fischer G, Wefer G (eds) (1999) Use of proxies in paleoceanography.

Springer, Berlin, p 735

Flint RF (1971) Glacial and quaternary geology. Wiley, New York, p

892

Flohn H (1985) Das Problem der Klimaa¨nderungen in Vergangenheit und Zukunft. Wissenschaftliche Buchgesellschaft, Darmstadt, p

228

Ganeshram RS, Pedersen TF (1998) Glacial-interglacial variability in upwelling and bioproductivity off NW Mexico: implications for quaternary paleoclimate. Paleoceanography 13(6):634–645

Ganopolski A, Rahmstorf S (2001) Rapid changes of glacial climate simulated in a coupled climate model. Nature 409:153–158

Gherardi JM et al (2005) Evidence from the North Eastern Atlantic

Basin for variability of the meridional overturning circulation through the last deglaciation. Earth Planet Sci Lett 240:

710–723

123

Int J Earth Sci (Geol Rundsch) (2010) 99 (Suppl 1):S171–S189 S187

Grinsted A, Moore J, Jevrejeva S (2009) Reconstructing sea level from paleo and projected temperatures 200 to 2100 AD. Clim

Dyn. doi: 10.1007/s00382-00008-00507-00382

Grootes PM, Stuiver M, White JWC, Johnsen S, Jouzel J (1993)

Comparison of oxygen isotope records from the GISP2 and

GRIP Greenland ice cores. Nature 366:552–554

Hansen J, Sato M, Kharecha P, Beerling D, Berner R, Masson-

Delmotte V, Pagani M, Raymo M, Royer DL, Zachos JC (2008)

Target atmospheric CO

2

: where should humanity aim? Open

Atmos Sci J 2:217–231

Hay WW, Brock JC (1992) Temporal variation in intensity of upwelling off Southwest Africa. In: Summerhayes CP, Prell WL,

Emeis KC (eds) Upwelling systems: evolution since the early miocene. Geol Soc (London) Spec Publ 64:463–497

Heinrich H (1988) Origin and consequences of cyclic ice rafting in the Northeast Atlantic Ocean during the past 130, 000 years.

Quat Res 29:143–152

Hurrell JW, Kushnir Y, Ottersen G, Visbeck M (2003) An overview of the North Atlantic Oscillation. In: Hurrell JW, Kushnir Y,

Ottersen G, Visbeck M (eds) The North Atlantic oscillation, climatic significance and environmental impact. AGU, Geophys,

Monograph 134:1–35

Imbrie J, Imbrie KP (1979) Ice ages: solving the mystery. Enslow,

Short Hills, NJ, 224 pp

Imbrie J, Kipp NG (1971) A new micropaleontological method for quantitative paleoclimatology: application to a late Pleistocene

Caribbean core. In: Turekian KK (ed) Late Cenozoic glacial ages. Yale University Press, New Haven, pp 71–181

Imbrie J, Hays JD, Martinson DG, McIntyre A, Mix AC, Morley JJ,

Pisias NG, Prell WL, Shackleton NJ (1984) The orbital theory of

Pleistocene climate: support from a revised chronology of the marine d 18O record. In: Berger AL, Imbrie J, Hays J, Kukla G,

Saltzman B (eds) Milankovitch and climate, vol 1. Reidel,

Hingham, MA, pp 269–305

IPCC (2007) Climate change 2007: the physical science basis.

Contribution of working group I to the fourth assessment report of the intergovernmental panel on climate change. Cambridge

University Press, Cambridge

Jackson SC, Broccoli AJ (2003) Orbital forcing of Arctic climate: mechanisms of climate response and implications for continental glaciation. Clim Dyn 21:539–557

Jansen E, Veum T (1990) Evidence for two step deglaciation and its impact on North Atlantic deep-water circulation. Nature

343:612–616

Jones PD, Osborn TJ, Briffa KR, Folland CK, Horton EB, Alexander

LV, Parker DE, Rayner NA (2001) Adjusting for sampling density in grid box land and ocean surface temperature time series. J Geophys Res 106:3371–3380

Kopp RE, Simons FJ, Mitrovica JX, Maloof AC, Oppenheimer M

(2009) Probabilistic assessment of sea level during the last interglacial stage. Nature 462:863–867

Ko¨ppen W, Wegener A (1924) Die Klimate der geologischen Vorzeit.

Borntra¨ger, Berlin

Ku¨nzi K (2002) Ozeane aus der Ferne gesehen - Die Fernerkundung als neues Werkzeug des Ozeanographen. Die Wittheit zu

Bremen - Der Ozean - Lebensraum und Klimasteuerung.

Hauschild, Bremen

Levitus S, Antonov JI, Wang J, Delworth TL, Dixon KW, Broccoli

AJ (2001) Anthropogenic warming of Earth’s climate system.

Science 292:267–270

Lisiecki LE, Raymo ME (2005) A Plio-Pleistocene stack of 57 globally distributed benthic d 18O records. Paleoceanography

20:PA1003. doi: 1010.1029/2004PA0001071

Lisiecki L, Raymo ME (2007) Plio-Pleistocene climate evolution: trends in obliquity and precession responses and glacial cycle shapes. Quat Sci Rev 26:56

Lototskaya A, Ganssen GM (1999) The structure of Termination II

(penultimate deglaciation and Eemian) in the North Atlantic.

Quat Sci Rev 18:1641–1654

Loulergue L, Schilt A, Spahni R, Masson-Delmotte V, Blunier T,

Lemieux B, Barnola JM, Raynaud D, Stocker TF, Chappellaz J

(2008) Orbital and millennial-scale features of atmospheric CH4 over the past 800, 000 years. Nature 453:383–386

Lu¨thi D, Le Floch M, Bereiter B, Blunier T, Barnola JM, Siegenthaler

U, Raynaud D, Jouzel J, Fischer H, Kawamura K, Stocker TF

(2008) High-resolution carbon dioxide concentration record 650,

000–800, 000 years before present. Nature 453:379–382

Manabe S, Stouffer RJ (1995) Simulation of abrupt climate change induced by freshwater input to the North Atlantic Ocean. Nature

378:165–167

MARGO Project Members, Waelbroeck C, Paul A, Kucera M,

Rosell-Mele´ A, Weinelt M, Schneider R, Mix AC, Abelmann A,

Armand L, Bard E, Barker S, Barrows TT, Benway H, Cacho I,

Chen MT, Cortijo E, Crosta X, de Vernal A, Dokken T, Duprat J,

Elderfield H, Eynaud F, Gersonde G, Hayes A, Henry M,

Hillaire-Marcel C, Huang CC, Jansen E, Juggins S, Kallel N,

Kiefer T, Kienast M, Labeyrie L, Leclaire H, Londeix L, Mangin

S, Matthiessen J, Marret F, Meland M, Morey AE, Mulitza S,

Pflaumann U, Pisias NG, Radi T, Rochon A, Rohling EJ, Sbaffi

L, Scha¨fer-Neth C, Solignac S, Spero H, Tachikawa K, Turon JL

(2009) Constraints on the magnitude and patterns of ocean cooling at the Last Glacial Maximum. Nat Geosci 2(2):127–132

Marotzke J (2000) Abrupt climate change and thermohaline circulation: Mechanisms and predictability. Proc Nat Acad Sci

97:1347–1350

Martin JH, Fitzwater SE (1988) Iron deficiency limits phytoplankton growth in north-east Pacific subarctic. Nature 331:341

Matthews RK (1990) Quaternary sea-level change. In: Revelle RR et al (eds) Sea-level change. National Academy Press, Washington, DC, pp 88–103

McGregor HV, Dima M, Fischer HW, Mulitza S (2007) Rapid 20thcentury increase in coastal upwelling off Northwest Africa.

Science 315:637–639

McManus JF, Francois R, Gherardi J-M, Kelgwin LD, Brown-Leger S

(2004) Collapse and rapid resumption of Atlantic meridional circulation linked to deglacial climate changes. Nature 428:834–837

Milankovitch M (1930) Mathematische Klimalehre und astonomische

Theorie der Klimaschwankungen. Handbuch der Klimatologie,

Bd 1, Teil A. Borntra¨ger, Berlin 176

Mix AC, Pisias NG, Rugh W, Wilson J, Morey A, Hagelberg TK

(1995) Benthic foraminifer stable isotope record from Site 849

(0-5 Ma): local and global climate changes. In: Pisias NG, Mayer

LA, Janecek TR, Palmer-Julson A, van Andel TH (eds) Proc.

ODP, Sci Results. College Station, TX 138:371–412

Monnin E et al (2001) Atmospheric CO

2 concentrations over the last glacial termination. Science 291(5501):112–114

Mudelsee M, Schulz M (1997) The Mid-Pleistocene climate transition: onset of 100 ka cycle lags ice volume build-up by 280 ka.

Earth Planet Sci Lett 151:117–123

Muller RA, MacDonald GJ (1997) Glacial cycles and astronomical forcing. Science 277:215–218

Oberha¨nsli H (1991) Upwelling signals at the northeastern Walvis

Ridge during the past 500,000 years. Paleoceanography 6:53–71

Oerlemans J, van der Veen CJ (1984) Ice sheets and climate. Reidel,

Dordrecht, p 217

Olausson E (1961) Sediment cores from the North Atlantic Ocean.

Rep Swed Deep-Sea Exped 7:227

Olausson E (1965) Evidence of climatic changes in North Atlantic deep-sea cores, with remarks on isotopic paleotemperature analysis. Progr Oceanogr 3:221–252

Oppenheimer M, Alley RB (2004) The West Antarctic ice sheet and long term climate policy. Climatic Change 64:1–10

123

S188 Int J Earth Sci (Geol Rundsch) (2010) 99 (Suppl 1):S171–S189

Otto-Bliesner BL, Schneider R, Brady EC, Kucera M, Abe-Ouchi A,

Bard E, Braconnot P, Crucifix M, Hewitt CD, Kageyama M,

Marti O, Paul A, Rosell-Mele A, Waelbroeck C, Weber SL,

Weinelt M, Yu Y (2009) A comparison of PMIP2 model simulations and the MARGO proxy reconstruction for tropical sea surface temperatures at last glacial maximum. Clim Dyn

32:799–815

Parker FL (1958) Eastern mediterranean foraminifera. Repts Swedish

Deep-Sea Exped 1947–1948, 8(4):219–283

Paul A, Berger WH (1997) Modellierung der Eiszeiten: Klimazyklen und Klimau¨berga¨nge. Geowissenschaften 15:20–27

Paul A, Scha¨fer-Neth C (2003) The Atlantic Ocean at the Last Glacial

Maximum: 2. Reconstructing the current systems with a global ocean model. In: Wefer G, Mulitza S, Ratmeyer V (eds) The

South Atlantic in the late quaternary: reconstruction of material budgets and current systems. Springer, Berlin, pp 549–583

Peltier WR (2002) On eustatic sea level history: last glacial maximum to Holocene. Quat Sci Rev 21(1–3):377–396

Petit JR, Jouzel J, Raynaud D, Barkov NI, Barnola J-M, Basile I,

Bender M, Chappellaz J, Davis M, Delaygue G, Delmotte M,

Kotlyakov VM, Legrand M, Lipenkov VY, Lorius C, Pe´pin L,

Ritz C, Satzman E, Stievenard M (1999) Climate and atmospheric history of the past 420, 000 years from the Vostok ice core, Antarctica. Nature 399:429–436

Pettersson H (1953) Westward ho with the albatross. Dutton & Co.,

New York, p 218

Phleger FB, Parker FL, Peirson JF (1953) North Atlantic Foraminifera, Sediment cores from the North Atlantic Ocean, Swedish

Deep-Sea Exped Repts 1947–1948, 7(1):3–122

Ragueneau O, Treguer P, Leynaert A, Anderson RF, Brzezinski MA,

DeMaster DJ, Dugdale RC, Dymond J, Fischer G, Franc¸ois R,

Heinze C, Maier-Reimer E, Martin-Jezequel V, Nelson DM,

Queguiner B (2000) A review of the Si cycle in the modern ocean: recent progress and missing gaps in the application of biogenic opal as a paleoproductivity proxy. Global Planet

Change 26:317–365

Rahmstorf S (2001) Abrupt climate change. In: Steele J, Thorpe S,

Turekian K (eds) Encyclopedia of ocean sciences, vol 1.

Academic Press, London, pp 1–6

Rahmstorf S (2007) A semi-empirical approach to projecting future sea-level rise. Science 315:368–370

Raper SC, Braithwaite RJ (2005) Low sea level rise projections from the mountain glaciers and icecaps under global warming. Nature

439:311–313

Raymo ME (1997) The timing of major climate terminations.

Paleoceanography 12:577–585

Revelle R, Suess HE (1957) Carbon dioxide exchange between atmosphere and ocean and the question of an increase of atmospheric CO

2 during the past decades. Tellus 9:18–27

Revelle RR, Barnett TP, Barron EJ et al. (1990) Sea-level change.

Studies in geophysics. National Research Council. National

Academy Press, Washington, DC, p 234

Rohling EJ, Grant K, Hemleben C, Siddall M, Hoogakker BAA,

Bolshaw M, Kucera M (2008) High rates of sea-level rise during the last interglacial period. Nature Geoscience 1:38–42

Ruddiman WF, Raymo ME, McIntyre A (1986) Matuyama 41, 000year cycles: North Atlantic Ocean and northern hemisphere ice sheets. Earth Planet Sci Lett 80:117–129

Ruddiman WF, Raymo ME, Martinson DG, Clement BM, Backman J

(1989) Mid-Pleistocene evolution of Northern Hemisphere climate. Paleoceanography 4:353–412

Ryther JH (1969) Photosynthesis and fish production in the sea.

Science 166:72–76

Sarmiento JL, Gruber N, Brzezinski MA, Dunne JP (2004) Highlatitude controls of thermocline nutrients and low latitude biological productivity. Nature 427:56–60

Sarnthein M, Tiedemann R (1990) Younger Dryas-style cooling events at glacial terminations I-VI at ODP Site 658; associated benthic d 13

C anomalies constrain meltwater hypothesis. Paleoceanography 5:1041–1055

Sarnthein M, Stattegger K, Dreger D, Erlenkeuser H, Grootes P,

Haupt BJ, Jung S, Kiefer T, Kuhnt W, Pflaumann U, Scha¨fer-

Neth C, Schulz H, Schulz M, Seidov D, Simstich J, van Kreveld

S, Vogelsang E, Vo¨lker A, Weinelt M (2001) Fundamental modes and abrupt changes in North Atlantic circulation and climate over the last 60 ky—concepts, reconstructions and numerical modeling. In: Scha¨fer P, Ritzrau W, Schlu¨ter M,

Thiede J (eds) The Northern North Atlantic: a changing environment. Springer, Berlin, pp 365–410

Sarnthein M, Gersonde R, Niebler H-S, Pflaumann U, Spielhagen R,

Thiede J, Wefer G, Weinelt M (2003) Overview of glacial

Atlantic Ocean mapping (GLAMAP-2000). Paleoceanography

18. doi: 10.1029/2002PA00769

Scha¨fer-Neth C, Paul A (2003) The Atlantic Ocean at the last glacial maximum: 1. Objective mapping of the GLAMAP seasurface conditions. In: Wefer G, Mulitza S, Ratmeyer V (eds) The South

Atlantic in the late quaternary: reconstruction of material budgets and current systems. Springer, Berlin, pp 531–548

Schlitzer R (1996) Mass and heat transports in the South Atlantic derived from historical hydrographic data. In: Wefer G, Berger

WH, Siedler G, Webb DJ (eds) The South Atlantic: present and past circulation. Springer, Berlin, pp 305–323

Schneider RR, Mu¨ller PJ, Ruhland G, Meinecke G, Schmidt H, Wefer

G (1996) Late quaternary surface temperatures and productivity in the east-equatorial South Atlantic: response to changes in trade/monsoon wind forcing and surface water advection.

In: Wefer G, Berger WH, Siedler G, Webb DJ (eds) The

South Atlantic: present and past circulation. Springer, Berlin, pp 527–551

Schott W (1935) Die Foraminiferen in dem a¨quatorialen Teil des

Atlantischen Ozeans. Deutsch Atlant Exped Meteor 1925–1927

Wiss Ergebn 3:43–134

Schulz M, Berger WH, Sarnthein M, Grootes PM (1999) Amplitude variations of 1470-year climate oscillations during the last 100,

000 years linked to fluctuations of continental ice mass. Geophys

Res Lett 26:3385–3388

Schulz M, Paul A, Timmermann A (2002) Relaxation oscillators in concert: a framework for climate change at millennial timescales during the late Pleistocene. Geophys Res Lett. 29: doi: 10.1029/

2002GL016144

Schwarzbach M (1974) Das Klima der Vorzeit, 3rd edn. Ferdinand

Enke, Stuttgart, p 380

Seager R, Kushmir Y, Visbeck M, Naik N, Miller J, Krahmann G,

Cullen H (2000) Causes of Atlantic Ocean climate variability between 1958 and 1998. J Climate 13:2845–2862

Seibold E, Berger WH (1996) The sea floor, 3rd edn. Springer,

Heidelberg, p 356

Shackleton NJ, Opdyke ND (1973) Oxygen isotope and paleomagnetic stratigraphy of equatorial Pacific core V 28–238: oxygen isotope temperatures and ice volumes on a 10^5 and 10^6 year scale. Quat Res 3:39–55

Shackleton NJ, Berger A, Peltier WR (1990) An alternative astronomical calibration of the lower Pleistocene timescale based on ODP Site 677. Trans R Soc Edinburgh, Earth Sci

81:251–261

Shepard FP (1963) Submarine geology. Harper and Row, New York, p 557

Siddall M, Rohling EJ, Almogi-Labin A, Hemleben C, Meischner D,

Schmelzer I, Smeed DA (2003) Sea-level fluctuations during the last glacial cycle. Nature 423:853–855

Siddall M, Bard E, Rohling EJ, Hemleben C (2006) Sea-level reversal during termination II. Geology 34:817–820

123

Int J Earth Sci (Geol Rundsch) (2010) 99 (Suppl 1):S171–S189 S189

Siegenthaler U, Stocker TF, Monnin E, Lu¨thi D, Schwander J,

Stauffer B, Raynaud D, Barnola JM, Fischer H, Masson-

Delmotte V, Jouzel J (2005) Stable carbon cycle-climate relationship during the late pleistocene. Science 310:1313–1317

Sigman DM, Boyle EA (2000) Glacial/interglacial variations in atmospheric carbon dioxide. Nature 407:859–869

Sima A, Paul A, Schulz M, Oerlemans J (2006) Modeling the oxygenisotopic composition of the North American ice sheet and its effect on the isotopic composition of the ocean during the last glacial cycle. Geophys Res Lett 33:LI5706

Soden BJ, Held IM (2006) An assessment of climate feedbacks in coupled ocean-atmosphere models. J Climate 19:3354–3360

Steffensen JP, Andersen KK, Bigler M, Clausen HB, Dahl-Jensen D,

Fischer H, Goto-Azuma K, Hansson M, Johnsen SJ, Jouzel J,

Masson-Delmotte V, Popp T, Rasmussen SO, Ro¨thlisberger R,

Ruth U, Stauffer B, Siggaard-Andersen ML, Sveinbjo¨rnsdottir

AE, Svensson A, White JWC (2008) High-resolution Greenland ice core data show abrupt climate change happens in few years.

Science 321:680–684

Stocker TF, Marchal O (2000) Abrupt climate change in the computer: is it real? Proceed National Acad Sciences 97:1362–

1365

Stommel H (1961) Thermohaline convection with two stable regimes of flow. Tellus 13:224–230

Stommel H (1966) The Gulf stream. University of California Press,

Berkeley, p 248

Stommel H (1980) Asymmetry of interoceanic fresh-water and heat fluxes. Proc Natl Acad Sci USA Geophys 77:2377–2381

Stroeve J, Holland MM, Meier W, Scambos T, Serreze M (2007)

Arctic sea ice decline: faster than forecast. Geophys Res Lett

34:L09501. doi: 09510.01029/02007GL029703

Suess E (1980) Particulate organic carbon flux in the oceans—surface productivity and oxygen utilization. Nature 288:260–263

Sundquist ET (1985) Geological perspectives on carbon dioxide and the carbon cycle. In: Sundquist ET, Broecker WS (eds) The carbon cycle and atmospheric CO

2

: natural variations archean to present. Geophysical Monograph 32. American Geophysical

Union, Washington, DC, pp 5–59

Sverdrup HU, Johnson MW, Fleming RH (1942) The oceans, their physics, chemistry, and general biology. Prentice-Hall, Englewood Cliffs, NY, p 1087

Tarasov L, Peltier WR (2005) Arctic freshwater forcing of the

Younger Dryas cold reversal. Nature 435(7042):662–665

Thompson LG et al (2004) Low-latitude mountain glacier evidence for abrupt climate changes. Eos Trans AGU 85 (47), Fall Meet

Suppl, Abstract C44A–02 (Invited)

Usbeck R, Schlitzer R, Fischer G, Wefer G (2003) Particle fluxes in the ocean: comparison of sediment trap data with results from inverse modeling. J Marine Systems 39:167–183

Van der Veen CJ (ed) (1999) Fundamentals of glacier dynamics.

Balkema, Rotterdam, p 462 von Dobeneck T, Schmieder F (1999) Using rock magnetic proxy records for orbital tuning and extended time series analyses into the super- and sub-Milankovitch bands. In: Fischer G, Wefer G

(eds) Use of proxies in plaeoceanography: examples from the

South Atlantic. Springer, Berlin, pp 601–633

Waelbroeck C, Frank N, Jouzel J, Parrenin F, Masson-Delmotte V,

Genty D (2008) Transferring radiometric dating of the last interglacial sea level high stand to marine and ice core records.

Earth Planet Sci Lett 265:183–194

Wallace AR (1895) Island life or the phenomena and causes of insular faunas and floras including a revision and attempted solution of the problem of geological climates, 2nd edn. Macmillan,

London, p 563

Wang C, Picaut J (2004) Understanding ENSO physics—a review. In:

Wang C, Xie S-P, Carton JA (eds) Earth’s climate: the oceanatmosphere interaction. AGU Geophys Monograph 147:21–48

Warrick R, Oerlemans J (2001) Sea level rise. In: Houghton JT,

Jenkins GJ, Ephraums JJ et al (eds) Climate Change. The IPCC

Scientific Assessment. Cambridge University Press, Cambridge,

UK, pp 257–281

Wefer G (1989) Particle flux in the ocean: effects of episodic production. In: Berger WH, Smetacek VS, Wefer G (eds)

Productivity of the ocean: present and past. Wiley, Chichester, pp 139–154

Wefer G, Berger WH (1991) Isotope paleontology: growth and composition of extant calcareous species. Mar Geol 100:207–

248

Wefer G, Fischer G (1993) Seasonal patterns of vertical particle flux in equatorial and coastal upwelling areas of the eastern Atlantic.

Deep-Sea Res 40:1613–1645

Wefer G, Berger WH, Siedler G, Webb DJ (eds) (1996) The South

Atlantic: present and past circulation. Springer, Berlin, p 644

Wefer G, Berger WH, Richter C, Shipboard Scientific Party (1998)

Initial Reports, Leg 175, Benguela Current. Proceedings of the

Ocean Drilling Program, vol 175 (2 pts.), Texas A&M, College

Station

Wefer G, Berger WH, Bijma J, Fischer G (1999) Clues to ocean history: a brief overview of proxies. In: Fischer G, Wefer G (eds)

Use of Proxies in Paleoceanography: examples from the South

Atlantic. Springer, Berlin, pp 1–68

Wefer G, Berger WH, Behre K-E, Jansen E (eds) (2002) Climate development and history of the North Atlantic Realm. Springer,

Berlin, p 486

Wefer G, Lamy F, Mantoura F (eds) (2003a) Marine science frontiers for Europe. Springer, Berlin, p 302

Wefer G, Mulitza S, Ratmeyer V (eds) (2003b) The South Atlantic in the late quaternary: reconstruction of material budgets and current systems. Springer, Berlin, p 722

Zachos JC, Pagani M, Sloan L, Thomas E, Billups K (2001) Trends, rhythms, and aberrations in global climate 65 Ma to present.

Science 292:686–693

Zachos JC, Dickens GR, Zeebe RE (2008) An early Cenozoic perspective on greenhouse warming and carbon-cycle dynamics.

Nature 451:279–283

123

Related documents
Download