CO
PY
RI
GH
TE
D
MA
TE
RI
AL
I
Global contexts
C HA PTE R 1
Introduction
Sarah E. Metcalfe and David J. Nash
1.1 Why the tropics matter
1.1.1 Defining the tropics
In its strictest sense, the term ‘tropics’ refers to
those parts of the world that lie between the Tropic
of Cancer (23.4378 °N) and the Tropic of Capricorn
(23.4378 °S). These latitudinal boundaries mark,
respectively, the most northerly and southerly
position at which the Sun may appear directly
overhead at its zenith. Indeed, the word ‘tropical’
comes from the Greek tropikos, meaning ‘turn’,
since the tropics of Cancer and Capricorn mark the
latitudes at which the Sun appears to turn in its
annual motion across the sky. Unfortunately, the
outer boundary of the tropics sensu lato cannot be
defined in such rigid astronomical terms. Certainly
latitude is a major factor determining the distribution of tropical climatic zones, through its control
on solar radiation receipt (Fig. 1.1), but regions
with distinctive climatological, physical or biological characteristics are not easily delimited by linear
boundaries.
The tropics include a diverse range of environments and climates (see Chapter 2). Rather than
being uniformly hot and wet, the area between the
tropics of Cancer and Capricorn encompasses some
of the wettest regions on Earth (e.g. the rainforests
of western Amazon and central Congo basins) as
well as some of the driest (e.g. the Atacama Desert
of northern Chile and Peru). The one feature
common to all tropical climates is a relatively
limited seasonal fluctuation in insolation and tem-
perature. Instead, differences in the quantity and
temporal distribution of available moisture account
for regional and seasonal variability (Savage et al.,
1982).
Authors such as Reading et al. (1995) have provided useful overviews of the various attempts to
define the climates of the tropics. Some of the most
widely used classifications are based directly upon
meteorological parameters such as rainfall and
temperature. The classic Köppen–Geiger system
(Fig. 1.2), for example, centres on the concept
that natural vegetation is the best expression of
climate, with climate zone boundaries positioned
with vegetation distribution in mind. The Köppen–
Geiger scheme combines average annual and
monthly temperatures and precipitation, and the
seasonality of precipitation. Köppen (1936) defined
tropical climates as those exhibiting a constant
high temperature (at sea level and low elevations),
with all 12 months of the year having average
temperatures of 18 °C or higher. This classification
excludes cooler highland regions (defined as areas
above 900 m elevation), which comprise around
25% of the total land area within the tropics
(Reading et al., 1995). These regions still receive
high amounts of solar radiation and do not have
a pronounced winter season, but temperatures
may be sufficiently depressed to affect biological
activity. Rainfall levels and the seasonal distribution of precipitation are then used to subdivide
tropical climates into tropical rainforest (Af), tropical monsoon (Am), and tropical savanna climates
Quaternary Environmental Change in the Tropics, First Edition. Edited by Sarah E. Metcalfe and David J. Nash.
© 2012 John Wiley & Sons, Ltd. Published 2012 by John Wiley & Sons, Ltd.
3
4
Chapter 1
20
20°S
30°N
20°S
20
Megajoules/day
10°N
15
10
0°N
15
0°S
10°N
20°S
10
30°N
5
0
5
30°N
J
60°N
F
M
60°N
45°S
A
M
J
J
A
S
O
N
D
0
Fig. 1.1 Solar radiation received at the Earth’s surface assuming an atmospheric transmission coefficient of 0.60
(after McGregor and Nieuwolt (1998) Tropical Climatology, John Wiley & Sons Ltd.).
(Aw). Köppen (1936) includes a range of other
climate types within the tropics sensu stricto, including tropical and subtropical steppe (BSh), tropical
desert (BWh) and humid subtropical climates (Cfa,
Cwa). Some highland areas within the tropics also
exhibit a temperate climate with dry winters (Cwb).
Working from an agricultural perspective,
Jackson (1989) split the tropics into three zones
(Humid, Wet and Dry, and Dry) according to the
level and seasonal distribution of rainfall (Fig. 1.3).
This classification recognises the importance of
seasonality for agricultural productivity, and is
less focused on natural vegetation zones than the
Köppen–Geiger scheme. Other attempts to classify
climates within the tropics are based around
hydro-meteorology, with climate types defined
according to the balance of precipitation inputs
and evapotranspiration outputs. Garnier (1958),
for example, differentiated humid tropical
climates according to the number of months in
which actual evapotranspiration equals potential
evapotranspiration. The ratio of precipitation to
potential evapotranspiration has also been used
by Middleton et al. (1997), drawing upon Thornthwaite (1948) and Meigs (1953), to define an
aridity index for categorising dry tropical climates.
In this volume, the astronomical definition of
the tropics is used to broadly demarcate the geographical scope of each of the substantive chapters.
However, recognising that climate boundaries are
fuzzy and mobile in the present day and that
climate zones shifted by many degrees of latitude
during the various glacials and stadials that characterise the Quaternary Period, coverage in many
chapters extends polewards north and south of
23.4378° into the subtropics where appropriate.
The Quaternary Period is defined here as encompassing the last 2.58 million years of the Earth’s
history (Gibbard et al., 2010), the timescale ratified
by the Executive Committee of the International
Union of Geological Sciences in June 2009.
1.1.2 Importance of the tropics
In comparison with the mid latitude regions of
Europe and North America, our understanding of
Quaternary palaeoenvironments in the tropics
is, at best, patchy for some areas and extremely
poor to non-existent in others. As a result, any
attempt to expand our understanding of past
environmental conditions in low latitude regions is
likely to be a valuable contribution to knowledge.
However, more significantly, understanding tropical
palaeoenvironments may also be key to establishing the drivers of global environmental change. As
discussed in section 1.5 of this chapter, the last
10–15 years have seen an increasing recognition
of the significance of tropical regions in climate
forcing (e.g. Kerr, 2001; Broecker, 2003). The
tropical oceans and atmosphere play an important
contemporary role in redistributing incoming solar
radiation and would have been instrumental in
transmitting past variations in radiation receipt to
Fig. 1.2 The Köppen–Geiger climate classification system updated with CRU TS 2.1 temperature and VASClimO v1.1 precipitation data for 1951 to 2000
(after Kottek et al., 2006). (See Colour Plate 1)
Introduction
5
6
Chapter 1
Fig. 1.3 Classification of the tropics based on the seasonal distribution of rainfall (after Jackson (1989) Climate, water and agriculture in the tropics,
Longman; Reading et al. (1995) Humid tropical environments, John Wiley & Sons Ltd.) (See Colour Plate 2)
Introduction
other parts of the Earth system. Tropical oceans and
landmasses also act as sources and sinks of greenhouse gases, with, for example, tropical forests
acting as contemporary CO2 sinks (Cox et al., 2000)
and tropical oceans (IPCC, 2007) and major river
and wetland systems such as the Amazon (Richey
et al., 2002) outgassing CO2 to the atmosphere. The
decay of vegetation within tropical wetlands is a
major source of contemporary biogenic CH4 (Loulergue et al., 2008). Indeed, much of the variation
in CH4 concentration recorded in the Antarctic
Vostok ice core coincides with fluctuations in the
size and extent of tropical lakes and wetlands
(cf. Raynaud et al., 1988; Chappellaz et al., 1990;
Brook et al., 2000). Tropical forest ecosystems and
soils are known to act as important contemporary
sources for atmospheric N2O, with N2O emissions
typically increasing during wet season conditions
and falling during drier periods. Data from the
Antarctic EPICA Dome C ice coring site suggest
that biospheric changes in the low latitudes may
have been instrumental in controlling emissions of
N2O on glacial–interglacial timescales (Schilt et al.,
2010). The precise mechanism through which
this process operated is unknown, but deep water
changes in the North Atlantic, and associated
Dansgaard–Oeschger (D–O) events, may have had
an influence on atmospheric N2O levels, either
through indirect changes in low latitude ecosystems and soils or by a direct change in marine N2O
production (Schmittner and Galbraith, 2008).
Identifying changes in tropical environments
over the past 2–3 million years may have considerable resonance for our understanding of the drivers
of human evolution. Recent fossil discoveries and
advances in the analysis of existing fossil collections, coupled with the emergence of high resolution palaeoclimatic records, have focused attention
on the role that past shifts in climate variability
may have had in the evolutionary history of African
mammalian fauna, including early hominids (de
Menocal, 2004). Although this topic is still hotly
debated, the basic premise is that large-scale shifts
in climate over the course of the last 5–6 million
years altered the ecological composition of African
landscapes, thereby generating specific faunal
adaptation or speciation pressures which ultimately
7
led to genetic selection and innovation. Evidence
from Atlantic and Indian Ocean cores suggests that
climate in the African subtropics fluctuated between
markedly wetter and drier conditions in time with
orbital variations. De Menocal (1995) identifies
progressive shifts in African climate variability
and increasing aridity after 3.0–2.6 Myr, 1.8–1.6 Myr
and 1.2–0.8 Myr, coincident with the onset and
intensification of high-latitude glacial cycles. Analysis of well-dated mammal fossil databases suggests
African faunal assemblage and, perhaps, speciation
changes coincident with the appearance of more
varied and open habitats at 2.9–2.4 Myr and after
1.8 Myr. These periods roughly coincide with key
junctures in hominid evolution, including the
emergence of the genus Homo around 2.5 Myr (de
Menocal, 2004).
Environmental changes, particularly during the
late Pleistocene, may also have played a role in
shaping pathways for the dispersal of early modern
humans around the Earth. For example, corridors
formed by pluvial ‘mega-lakes’ during Marine
Isotope Stage (MIS) 5 may have provided transSaharan pathways for humans migrating ‘out of
Africa’, offering an alternative route to the Nile
Valley (Drake et al., 2011). Biogeographical and
palaeohydrological evidence (ibid.) suggests that
similar migration pathways across the Sahara, in
the form of linked lakes, rivers and inland deltas,
may have existed during the early Holocene (see
Chapter 4). The migration of humans into Australia, either as a single or several successive waves,
also appears to have been influenced by global
environmental changes. There is still much debate
about the timing of the earliest arrivals; the
minimum widely-accepted timeframe places this at
around 45 kyr BP (e.g. O’Connell and Allen, 2005)
with an upper estimate of around 60 kyr BP (e.g.
Roberts et al., 1990, 1993, 1994). Regardless, this
migration was achieved during the closing stages of
the Pleistocene, when sea levels were much lower
than they are today (see section 1.2.2 and Fig. 1.5)
and an extensive land bridge existed across the
Arafura Sea, Gulf of Carpentaria and Torres Strait
(Lourandos, 1997).
The tropics are also highly important because
they support very large numbers of species compared
8
Chapter 1
with other regions of the globe (Mace et al., 2005).
This is especially true of the tropical moist forests
which show the highest global levels of species
and family richness and of endemism. There is
increasing concern about the threat posed to
tropical ecosystems by both direct human action
and by future climate change (itself probably
anthropogenic). Although we hear most about the
tropical rainforest (e.g. Hubbell et al., 2008), it is
the tropical dry forests that have been most affected
to date, with about half being lost to cultivation.
Mapping of species loss (mammals, birds and
amphibians) since AD 1500 shows a significant
concentration in tropical latitudes, especially in the
tropical Americas and Australasia (Baillie et al.,
2004). As well as direct loss of habitat and species,
with their economic and cultural values, changes
in tropical ecosystems have wider implications
because of their role in the global biogeochemical
and hydrological cycles. Some of these issues are
discussed further in section 1.5 of this chapter.
As the chapters within this volume highlight,
tropical vegetation and landscape systems have
shown considerable sensitivity to climatic changes
during the Quaternary Period; by inference, tropical
landscapes might be expected to show a similar
scale of response to future human-induced and
natural environmental changes. Couplings between
vegetation cover and the susceptibility of the
ground surface to water or wind erosion mean that
shifts in vegetation density and type in response to
anthropogenic and climatic changes may act to
either stabilise or destabilise land surfaces. Tropical
fluvial systems, for example, are highly sensitive
to external forcings in the form of short and long
term shifts in effective precipitation and vegetation
cover. The nature of the response within individual
fluvial systems reflects the antecedent conditions,
the degree and duration of the environmental
change, possibly the rate of change, and whether
the change is sufficient to trigger in-channel
threshold-crossing events (Thomas, 2008). In
northeast Queensland, Australia, for example, a
long-term deterioration of the rainforest vegetation
cover after 78 kyr BP, steepening after 40 kyr BP
with a shift toward dry sclerophyll forest, led to
widespread soil erosion and the accumulation of
fine alluvial fan deposits within fluvial systems
fronting the eastern highlands of the Great Dividing
Range (Nott et al., 2001; Thomas et al., 2001,
2007). Many tropical environments contain relict
landforms (and their associated sediments) formed
under previously wetter or drier conditions, which
may be reactivated under future climatic change
scenarios. Environmental modelling studies in the
Kalahari Desert, for example, have suggested that
large areas of presently stable and well-vegetated
‘fossil’ Pleistocene sand dunes could be reactivated
if changes in wind regime and a reduction in
vegetation cover (in response to warming and reduced
available moisture) occur as a result of twentyfirst century climate warming (Thomas et al., 2005).
1.2 Development of ideas
1.2.1 Early ideas about tropical
environmental change
The possibility that high and mid latitude regions
had undergone major environmental changes
was recognised as early as 1779 when the Swiss
aristocrat Horace-Bénédict de Saussure identified
granite boulders on the limestone slopes of the Jura
ranges that had been transported some 90 km from
their source in the Mont Blanc massif (de Saussure,
1779). In keeping with contemporary ideas that the
Earth’s features had been shaped by the biblical
Great Flood, de Saussure suggested that these
‘erratics’ had been moved by water. Bernard
Friedrich Kuhn was the first to propose that the
boulders had, in fact, been transported by more
extensive glaciers (Kuhn, 1787; de Beer, 1953), a
conclusion reached independently some eight
years later by James Hutton following a visit to the
Jura (Hutton, 1795). John Playfair famously
extended these ideas in 1802, and, by the time of
the publication of Etudes sur les Glaciers by Louis
Agassiz in 1840, the concept of Die Eiszeit or large
scale Ice Age in Europe was well established.
In contrast, for many years, the dominant view
of the tropics was that they had seen very little
climatic change, with core areas such as Amazonia
remaining unaffected by the cycles of glaciation
and deglaciation that drove massive environmental
Introduction
changes in higher latitudes (Richards, 1952). This
was despite the suggestion made by Louis Agassiz,
after mistaking deeply weathered bedrock for
glacial diamicton during a visit to Brazil in 1865–
1866, that the western Amazon basin had been
glaciated (Agassiz, 1868). As early as 1850, the
Scottish missionary and explorer David Livingstone
had recognised that salt accumulations in the
Makgadikgadi Depression of Botswana were ‘the
remains of the very slightly brackish lakes of antiquity’ (Livingstone, 1857: 67). However, some of
the main advances in our understanding of low
latitude palaeoenvironments were made in the
USA (see Goudie, 1999). John Strong Newberry,
for example, suggested that the landscapes of the
Colorado Plateau were ‘formerly much better
watered than they are now’ (1861: 47). In 1863,
Thomas Francis Jamieson was the first to propose
that wetter conditions and higher lake levels in
the southwest USA were equated with high
latitude glacial episodes (a concept often termed
the ‘glacial = pluvial’ hypothesis). This idea was
adopted by Israel Russell (1885) and Grove Karl
Gilbert (1890) to explain the origins of strandlines
within the Pleistocene ‘pluvial’ lakes Lahontan and
Bonneville (Fig. 1.4). The notion that low latitude
pluvials were synchronous with high latitude glacials was widely accepted and was ultimately
assumed to apply across the tropics. The corollary
of this view, that post-glacial times were characterised by desiccation, was also widely applied (Goudie,
1972), most notably in the Asian and African
tropics and subtropics (Goudie, 1999). In southern
Africa, for example, Schwarz (1923) proposed a
grandiose scheme to divert rivers from the north to
flood the Kalahari Basin as a means to ameliorate
a supposed progressively drying climate.
(a)
(b)
1.2.2 The twentieth century revolution
By the mid 1940s, challenges to the post-glacial
desiccation and ‘glacial = pluvial’ hypotheses began
to emerge. One of the most important conceptual
advances was the recognition that some tropical
areas that are now relatively moist, far from progressively desiccating had been significantly drier
in the past. The main evidence for this came first
from the identification of ancient dunefields in
Fig. 1.4 (a) Sketch of Lake Bonneville shorelines and
(b) Map of Lake Bonneville by G.K. Gilbert (from
Gilbert, 1890, images courtesy of USGS).
9
10
Chapter 1
Texas (Price, 1944), and then vegetated ergs along
the equatorward margins of the southern Sahara
(Grove, 1958; Grove and Warren, 1968), northern
Kalahari (Grove, 1969) and Indian deserts (Goudie
et al., 1973). Prior to the advent of luminescence
dating in the 1980s, the ages of these aeolian
deposits could only be estimated relative to sediments that could be radiocarbon dated, but their
mere existence was a nail in the coffin for progressive desiccation.
The 1970s represented a major shift in our
understanding of low latitude palaeoenvironments,
as detailed records from lake basins in tropical
Africa (e.g. Grove and Goudie, 1971; Grove et al.,
1975; Street and Grove, 1976) and elsewhere (cf.
Street-Perrott et al., 1979) started to be published.
Views of the stability of the tropical rainforest
also changed (Flenley, 1979). With these studies, it
became apparent that the story of the tropics was
much more complex than previously thought, with
many areas exhibiting fluctuating rather than consistently high lake levels around the time of the
Last Glacial Maximum (LGM). Compilations of
global lake level fluctuations (e.g. Street-Perrott
et al., 1979) served to demonstrate that, in many
ways, what happened in the southwest USA, the
home of the ‘glacial = pluvial’ hypothesis, was the
exception rather than the norm. The picture that
emerges today, as summarised by each of the
regional chapters in this volume, is that the magnitude and timing of climate change in different
parts of the tropics and subtropics is considerably
more complex than pioneers such as Gilbert and
Russell ever could have envisaged.
In parallel with the growth in knowledge about
terrestrial tropical environments, our understanding of Quaternary stratigraphy in tropical oceans
has been revolutionised since the 1950s (Imbrie
and Imbrie, 1979). This has been due primarily to
the introduction of new equipment for coring offshore and deep-ocean sediments. Damuth and
Fairbridge (1970), for example, used evidence from
deep-sea piston cores taken in the Guiana Basin to
suggest that an arid to semi-arid climate dominated
large portions of equatorial South America during
the Pleistocene glacial phases. Similarly, analyses of
multiple cores off northwest Africa by Diester-Haas
(1976) revealed fluctuations in the extent of the
Sahara during the late Pleistocene. Some of the
longest and highest resolution records available for
the tropics now come from marine settings, and
provide important insights into ocean palaeotemperatures, terrestrial chemical environments and
variations in the offshore transport of dust, pollen
and fluvial sediments (e.g. Larrasoaña et al., 2003;
Peterson and Haug, 2006) (see Chapter 3, section
3.2.2). The technology used to extract marine cores
has been adapted and utilised on land, such that a
number of long terrestrial records are now also
available for the tropics (e.g. Trauth et al., 2003).
At the interface between land and the oceans,
the Quaternary has seen major changes in relative
sea level. A number of different factors may be
involved, especially locally. However, at the global
scale, glacio-eustatic change dominates, reflecting
the volume of water locked up in ice sheets
and glaciers (with glacial or stadial periods being
marked by low sea levels). Although the association between changes in ice volume and sea level
was put forward in the early twentieth century,
major advances in reconstructing sea level were
made during the 1960s and 1970s. Key sea level
reconstructions (covering about the last 400 kyr)
have come from tropical areas, primarily the Huon
Peninsula of Papua New Guinea (Aharon and
Chappell, 1986) (see Chapter 6) and Barbados
(Fairbanks, 1989). Both these are based on dated
sequences of coral reefs. An updated version of
Fairbanks’ reconstruction for the period since the
LGM is shown in Fig. 1.5. This indicates that sea
level was 120–125 m lower than present at the
LGM. Fairbanks identified two periods of very rapid
rise (>20 m) associated with meltwater pulses 1A
and B, which he dated to around 12 kyr and 9.5 kyr
BP. These events were later re-dated to 14 and
11 kyr BP, following a reassessment of the record
using U-Th dating (Bard et al., 1990). Bard et al.
also reported sea level of +5 to +10–m in the last
interglacial (MIS 5e). This and subsequent studies
have confirmed the coincidence of periods of high
sea level with insolation maxima, consistent
with Milankovitch forcing (see section 1.4 of this
chapter). The impact of these changes in sea level
was particularly pronounced in areas with exten-
Introduction
11
Fig. 1.5 Composite record of relative sea level change over the last 32 kyr, based on data from Barbados. Data from
Peltier and Fairbanks (2008) IGBP PAGES/World Data Center for Paleoclimatology, Data series 2008-101.
sive continental shelves affecting marine currents,
regional groundwater levels and the ease of migration of terrestrial organisms including humans (see
especially Chapters 6 and 7).
Advances in our understanding of tropical palaeoenvironments have been prompted, in part, by
the availability of new avenues for environmental
reconstruction, but also reflect the development of
new chronological techniques (Goudie, 1999). The
introduction of radiocarbon dating in the 1950s, for
example, meant that it was possible, for the first
time, to obtain age estimates from late Quaternary
sediments and landforms rather than having to
rely on stratigraphic correlation. The radiocarbon
revolution was followed in the 1960s by the development of potassium-argon and uranium-series
dating, dendrochronology and palaeomagnetism.
These chronological tools were refined in the 1970s
and 1980s, with new approaches such as amino-
stratigraphy, electron-spin resonance, luminescence
and cosmogenic radionuclide exposure dating introduced in more recent decades. For many of these
techniques, the availability of mass spectrometry
has permitted high temporal resolution dating of
materials, including the micro-sampling of cave
deposits (e.g. Wang X et al., 2007; Wang Y et al.,
2008) and geochemical sediments such as calcrete
(e.g. Candy et al., 2004).
Two examples serve to highlight the importance
of the new dating tools for our understanding
of tropical palaeoenvironments. First, the development of optically-stimulated luminescence (OSL)
dating since the 1980s has allowed the age of deposition of a wide range of carbon-poor sediments to
be determined, most notably those preserved
within fossil dunes and other aeolian deposits (cf.
Singhvi and Porat, 2008). This has led to the establishment of detailed chronological frameworks for
12
Chapter 1
many of the world’s desert regions (cf. Munyikwa,
2005) as well as major advances in our understanding of how aeolian dunes evolve over time
(e.g. Telfer and Thomas, 2006). Second, the rapid
advances in cosmogenic radionuclide analysis in
the last decade have provided a basis for exposure
‘dating′ of landforms, the quantification of erosion
rates and other geologic applications in areas where
opportunities for any form of chronological investigation were once extremely limited. Cosmogenic
radionuclide dating has been used, for example, to
establish the timing of dunefield initiation in central
Australia (Fujioka et al., 2005) and, alongside other
techniques, to estimate residence times for groundwater in the Nubian Aquifer beneath the Western
Desert in Egypt (Patterson et al., 2005).
Alongside chronological developments, there have
been a number of other methodological improvements, including the application of an increasingly
sophisticated range of field and laboratory approaches. These include new techniques for sedimentological and geochemical analysis which have
offered important insights into Quaternary depositional environments. Amongst the most significant
of these was the advent of stable isotope analyses
of sediments and biological remains in the 1950s.
Oxygen isotope analysis, in particular, pioneered by
Cesare Emiliani (1955), is now one of the most
important tools in Quaternary stratigraphy and is
routinely applied in a variety of terrestrial and
marine contexts to reconstruct environmental
signals such as palaeotemperature, water balance
(P–E), precipitation source and amount (Leng,
2006). Stable carbon isotope analysis can be
undertaken on either inorganic (authigenic calcite,
biogenic carbonate), or organic C. In combination
with measurements of C/N, δ13Corganic is widely used
to determine the sources (C3 or C4 terrestrial vegetation, aquatic macrophytes, algae) of organic
matter coming into lacustrine systems. The application of compound specific δ13C analysis is particularly effective in this regard (e.g. Street-Perrott
et al., 2004). More recent studies (e.g. Chase et al.,
2009), have utilised variations in stable nitrogen
isotope analyses as a means of establishing past
rainfall levels.
A large (and growing) number of palaeoecological techniques are now available for environmental
reconstruction. Most early attempts to reconstruct
changes in tropical flora were heavily reliant upon
pollen analyses. However, it is now possible to
utilise other plant remains such as macrofossils
(e.g. those preserved within rodent middens;
Betancourt et al., 1990; Pearson and Dodson, 1993;
Holmgren et al., 2007) and phytoliths (e.g. Parker
et al., 2004), not only to reconstruct terrestrial
vegetation changes but also to identify shifts in
CO2 concentration (Beerling and Woodward, 1993).
Changes in terrestrial aquatic environments can be
identified through the analysis of molluscs, diatoms
and ostracods (e.g. Fritz et al., 1999; Holmes
and Engstrom, 2005), whilst our understanding
of changes in marine environments has been
revolutionised through the analysis of foraminifera
and other microorganisms such as radiolaria
and coccoliths (cf. Lowe and Walker, 1997). The
development of transfer functions – essentially
variants on multiple linear regression models
employed to establish relationships between
biological data and environmental variables – now
permits palaeoenvironmental parameters to be
reconstructed quantitatively from fossil floral and
faunal assemblages (e.g. Birks and Birks, 1980;
Birks, 2005). The need for such transfer functions
to reflect biologically meaningful relationships has
to be borne in mind, however.
Finally, our understanding of tropical and subtropical environmental variability in recent centuries has greatly improved thanks to new efforts to
tap the wealth of information contained within
annually resolved proxies (e.g. corals, tree rings,
speleothems). Climate chronologies derived from
historical documentary materials are now available, for example, for large areas of Africa (e.g.
Nicholson, 2000, 2001; Nash and Endfield, 2002,
2008; Grab and Nash, 2010; Nash and Grab, 2010)
and show remarkable agreement with regional tree
ring records (e.g. Therrell et al., 2006) and fossil
coral (Zinke et al., 2004, 2005).
1.2.3 Advances in modelling
The application of computer modelling to palaeoclimate studies is now central to efforts to synthesise and understand change in climate systems and
environments. The role of factors such as insolation
forcing, tectonism and vegetation feedbacks have
Introduction
all been explored in relation to tropical regions,
with a particular emphasis on their impacts on
monsoons. The application of modelling to tropical
palaeoclimates is explored explicitly in Chapter 9,
so this section will provide only a brief introduction. The reader is also referred to a number of
reviews of climate modelling, including those of
McGuffie and Henderson Sellers (2001), Cane et al.
(2006) and the IPCC (2007).
Climate models are derived from weather forecasting models, originally conceived by John von
Neumann who founded the GFDL (Geophysical
Fluid Dyamics Laboratory). The first comprehensive general circulation experiments were undertaken by Smagorinsky (1963) and by 1965 it was
realised that computer models could also be used
to explore past climates. There are a range of model
types from 1-D energy balance models, to 3-D
general circulation models (GCMs). Pioneering
work on the application of modelling to palaeoclimate was carried out by Gates (1976a,b) and
Manabe and Hahn (1977). This work brought
climate modellers and palaeoclimatologists together,
as palaeodata (e.g. CO2 concentrations, sea-surface
temperatures (SSTs), ice sheet extents) were needed
to set model boundary conditions. Through the
1980s, John Kutzbach and his co-authors led the
way in exploring drivers of change in the monsoon
using the NCAR CCM (Community Climate Model)
(e.g. Kutzbach and Guetter, 1986; Prell and Kutzbach, 1987; Ruddiman and Kutzbach, 1989). This
effort was complemented by significant developments in data-model comparisons through
COHMAP with a particular focus on 18 k and 6 k
14
C yr BP (COHMAP Members, 1988). This tradition has been continued through the PMIP (Palaeoclimate Modelling Intercomparison Project).
PMIP1 used CGMs with atmosphere only, or with
slab ocean, while PMIP2 used coupled ocean–
atmosphere (–vegetation) models (Braconnot et al.,
2007). The results from PMIP are discussed in more
detail in Chapter 9.
The development of fully coupled ocean–
atmosphere models (see Chapter 9) represented a
major challenge due to the very different response
times and resolutions of these two key elements of
the climate system. Early coupled models such as
the UK Met Office’s HadCM2 required flux adjust-
13
ments to keep the two elements together, but
this wasn’t needed in later models. The advent
of these coupled models allowed annual climatologies and seasonal cycles to be reproduced (IPCC,
2001). This has been vital in efforts to model the
El Niño Southern Oscillation (ENSO; see Chapter
9). The most recent development is the use of fully
coupled Earth System Models such as HadGEM2ES (dynamic vegetation response) and ECHAM5/
JSBACH-MPIOM (e.g. Dallmeyer et al., 2010).
GCMs now dominate, but simpler models are still
used where long time series are a key requirement
(e.g. Crowley et al., 1992) and the run times of
more comprehensive models would still be prohibitive even with the significant computing
power now available. These models of intermediate
complexity continue to play an important role in
helping to understand long term climate change,
including the role of Milankovitch cycles and transitions between different climate modes (interglacial/
glacial). Groot et al. (2011) use one of these models,
CLIMBER (see also Chapter 9), to help interpret
the arboreal pollen record from the Fuquene Basin
in Colombia between 284 kyr and 27 kyr BP (see
Chapter 8).
Models play a very important part in helping our
understanding of tropical climate change. They
have also helped us to appreciate the importance
of the tropics in driving climate change, especially
the role of the tropical oceans (Hostetler et al.,
2006) and feedbacks from greenhouse gases (particularly methane) (Loulergue et al., 2008). Unfortunately, there are still some parts of the world
where climate models struggle to reproduce modern
climate, and hence one can have little confidence
in their use in palaeoclimatic studies. This is
particularly the case in areas of complex terrain.
The use of regional scale models and finer resolution GCMs can help to address this (e.g. Hostetler
et al., 1994).
1.3 Establishment of the tropical
climate system
In the popular imagination, the tropics are both
warm and wet, and it is the case that 56% of total
global precipitation falls in the tropics (Wang and
14
Chapter 1
Ding, 2008). As noted above, in tropical climates it
is the distribution of rainfall, rather than temperature, which determines the seasons, and the seasonality and overall amount of precipitation that
distinguishes the major tropical environments:
rainforest, savanna and desert (Bridgman and
Oliver, 2006). The reader is referred to Chapter 2
for more on tropical climatology, but in this section
some background is given on two key elements
of the tropical climate system: the monsoon
and ENSO.
Although the dominant dynamic controls on the
tropical climate are the location of the Intertropical
Convergence Zone (ITCZ) and the subtropical
high pressure systems (Hadley cells), perhaps the
best known feature of the tropical climate is the
monsoon. The name comes from the Arabic word
‘mausim’ for a seasonal reversal of winds recognised in the Arabian Sea and Indian Ocean and
exploited by Greek and Arab traders. The importance of this seasonal change in winds and the
resulting precipitation to trade (sailing ships) and
livelihoods (crops etc.) was recognised early on.
Failure of the monsoon rains in 1866 and 1871 led
to the establishment of the India Meteorological
Department in 1875 and the subsequent work of
H. F. Blanford and Sir G.T. Walker to forecast and
understand monsoon variability. In Walker’s case,
his analysis of meteorological data from around the
globe led to the recognition of the Southern Oscillation (identifying the importance of change in the
eastern tropical Pacific) and its link to monsoon
rainfall (Walker, 1924). The Southern Oscillation is
discussed further below.
Although there are various delineations of
monsoon areas – Wang and Ding (2008) suggest
that they cover 19.4% of the Earth’s surface –
monsoon rain accounts for 30.8% of total global
precipitation. Given that these areas are home to
more than 55% of the world’s human population
(McGregor and Niewolt, 1998) and support the
world’s most biologically diverse and ecologically
complex terrestrial ecosystems (tropical forests)
(Wilson, 1986) their significance is evident and
changes in monsoon climates both in the past and
into the future are important to understand. As
described in Chapter 2 (and Chapters 4 to 8 for
regional details), the monsoon climate is characterised by a reversal of prevailing wind direction and
a contrast between a wet summer and dry winter.
This seasonal reversal in wind direction (conventionally a change of ≥120° between January and
July) is driven by differential heating of oceans and
continents. Evaporation and condensation processes add strength to the system and Coriolis
results in the curved trajectories of monsoon winds.
Traditionally monsoons were associated with
Africa, Asia (India and East Asia) and Australia,
being best developed in South and Southeast Asia.
More recently monsoon-type systems have also
been identified in the tropical Americas, although
not fulfilling all the original criteria (McGregor
and Nieuwolt, 1998). In this volume, we adopt this
wider definition of monsoons.
It is clear that monsoons are an enduring feature
of the Earth’s climate system, with monsoon climates recognised in deposits from ancient super
continents (e.g. Pangaea) (Clift and Plumb, 2008).
It seems likely that the inception of the modern
Asian monsoon dates to the construction of Asia,
as it now exists, through the collision of the Indian
and Asian blocks around 45 to 50 Myr. The elevation of the Tibetan Plateau/Himalayas also appears
to be important, and early work on the effects of
mountains on monsoons was carried out by Hahn
and Manabe (1975). Prell and Kutzbach (1992)
linked the modern elevation of the Tibetan Plateau
to the strength of the monsoon. The date that Tibet
reached its present height is not clear (and may be
regionally variable). Around 8 Myr has been suggested, but estimates vary between 35 to less than
7 Myr and a high plateau may have existed before
8 Myr. There is evidence for stronger monsoons
after 8 Myr from deep sea cores in the Arabian Sea
(Kroon et al., 1991) and in Chinese loess sequences
which date back to 7–8 Myr. Loess itself is a proxy
for the winter monsoon, and the interbedded palaeosols for the summer monsoon. Loess–palaeosol
sequences may date back to more than 7 Myr (An,
2000) (see Chapter 6, section 6.2), but with a significant increase in loess accumulation since about
2.7 Myr. Harris (2006) questions whether the shift
around 8 Myr is actually due to the monsoon itself
or wider oceanographic changes associated with
Introduction
increasing glaciations of Antarctica. Elsewhere,
uplift in western North America is also seen as
important; Tibet and the Rockies reach high enough
elevations to disrupt the circulation in the upper
atmosphere, affecting the mean location and amplitude of winter planetary waves and the location of
the Siberian High.
Harris (2006) suggests that monsoon intensification may actually date from the Miocene–Oligocene
boundary (∼24–22 Myr) (based on data from the
South China Sea), with the East Asian monsoon
starting earlier than the Indian/Arabian monsoon.
The former is more dependent on the evolution
of the West Pacific Warm Pool and the latter on
the uplift of the Himalayas/Tibet. Harris highlights
the importance of tectonic influences on ocean
currents, particularly the severance of Indonesian
through flow and the closure of the Panama gateway
helping to create the modern Pacific Ocean.
The association between the development of the
modern monsoon and the onset of the last glacial
is a topic that has been widely debated. Ruddiman
and Raymo (1988) suggest that uplift in Tibet
and western North America played a role in the
intensification of glaciations over the Pliocene
that culminated in large scale glaciations about
2.4 Myr. The impact of uplift in these areas on CO2
drawdown (via the weathering effect) has also
been the focus of considerable interest. Raymo and
Ruddiman (1992) propose that uplift of the Tibetan
plateau over the last 40 Myr and the associated
increase in erosion, lowered global CO2, driving a
positive feedback of global cooling. Mudelsee and
Raymo (2005) provide a wider view of tectonic
forcing and the development of Northern
Hemisphere (NH) glaciations. The patterns of global
climatic, tectonic and biotic events are summarised
in Fig. 1.6. The theme of changes in the tropics
driving environmental change is explored further
below.
At interannual timescales, ENSO is the dominant
source of climatic variability in the tropics, where
the Southern Oscillation Index (or SOI) is a measure
of the strength of the Walker circulation (the east–
west circulation across the Pacific which in ‘normal’
years gives high pressure and dry conditions in the
east and low pressure and rain in the west). Major
15
weakening of the Walker circulation (low index
conditions) results in the warm phase of the SOI,
with warming of the eastern Pacific, weakening of
the easterly trade winds and the ITCZ south of its
usual position over South America. A deeper than
normal thermocline develops on the east side of
the Pacific, leading to a breakdown of the normal
upwelling current and the area of high precipitation
effectively moves east across the Pacific. The most
obvious impacts occur along the west coast
of tropical South America, with fisheries declining
and major increases in precipitation (and erosion).
The biggest impacts occur around Christmas, hence
the name El Niño (the boy child). ENSO displays
a cyclicity of 2–7 years and the strongest El Niño’s
of twentieth century occurred in 1982–1983 and
1997–1998. Changes originating in the tropical
Pacific have clear impacts on temperature and
precipitation around the world, including mid
latitudes (with seasonality) (Diaz and Markgraf,
2000). Significantly for tropical regions, El Niños
(the warm phase of the SOI) are generally
associated with weakened monsoons, and La Niñas
(the cold phase of the SOI) with strong monsoons.
As described above, it was Sir Gilbert Walker’s
efforts (as Director General of Observatories, India
Meteorological Department) to understand and
forecast monsoons that led him to identify the
Southern Oscillation, based on changes in the
pressure gradient between Tahiti and Darwin.
The global reach of the effects of El Niño have
made understanding both its past and possible
future a major research focus and the severe El Niño
of 1982–1983 was a further stimulus to research.
There have been a number of major syntheses of
ENSO and its impacts including those of Diaz and
Markgraf (1992, 2000) and Sarachik and Cane
(2010). These draw on a range of sources including
the (relatively short) instrumental record, historical
records and proxy data. The first major publication
based on historical records was that of Quinn et al.
(1987), spanning 450 years, and focusing on Peru
and southern Ecuador. Since then, there have been
a number of syntheses of historical records including Ortlieb (2000) and Gergis and Fowler (2009).
Records for El Niño per se extend back to the last
interglacial, based on corals (Tudhope et al. 2001).
16
Chapter 1
r
v
Fig. 1.6 Summary of global environmental changes over the last 65 Myr. Redrawn from Zachos, J., Pagani, M.,
Sloan, L., et al. (2001) Trends, rhythms and aberrations in global climate 65Ma to present, Science 292, 686–93,
copyright (2001), with permission from the American Association for the Advancement of Science.
Other sources of palaeo-ENSO records include
some lake and marine sediments, ice cores and tree
rings (see Chapter 8, section 8.7 in particular).
These provide evidence of changes in both frequency and intensity in ENSO (weaker in glacials,
stronger in interglacials). There have evidently also
been changes within the Holocene, with the ENSO
being weaker during the early–mid Holocene and
stronger after about 5 kyr BP, possibly with a major
step up after 3 kyr BP. This variability in the
Holocene has been linked to orbital changes (see
also section 1.4), with early Holocene stronger
summer insolation in the Northern Hemisphere
(NH) resulting in stronger trade winds and mon-
soons and a weaker ENSO cycle. One question is
whether ENSO has ever stopped completely. Based
on modelling, Clement et al. (2001) suggest that it
did in the Younger Dryas (or YD; c. 13–11.5 kyr BP)
and possibly before.
In addition to ENSO itself there is increasing
interest in persistent El Niño-type states, also called
ENSO-type variability, over decades to millennia
(see for example Chapter 7). Molnar and Cane
(2002) suggest a virtually permanent El Niño like
state in the Pliocene (5 to 2.7 Ma), with strong
Hadley cell circulation driving more heat transport
into mid and high latitudes. Onset of this state may
have been driven by closure of the Indonesian
Introduction
seaway (see above). The authors suggest that this
El Niño-type state might explain a lack of major NH
ice sheets (winters warmer and drier). When this
situation came to an end, major NH glaciations set
in and cold upwelling zones became established off
the west coasts of Africa and America. Fedorov
et al. (2006) put this change into the context of
gradual cooling over the Cenozoic which eventually reduced the depth of the ocean thermocline
and allowed cold waters to come to the surface.
Increasingly detailed marine records from both the
normally cold eastern equatorial Pacific (Lawrence
et al., 2006) and West Pacific Warm Pool (MedinaElizalde and Lea, 2010) have explored the Pliocene
to Pleistocene transition and possible drivers,
emphasising the importance of the tropics.
Modelling of ENSO is discussed in some detail in
Chapter 9, section 9.4.1, but a recent review by
Sarachik and Cane (2010) sees this as still problematic, even with state of the art coupled models (e.g.
IPCC, 2007). These shortcomings with current
models cause problems for both ENSO forecasting
(short term) and modelling changes in ENSO over
the longer term (see Chapter 11, section 11.3.1).
Simpler coupled models (e.g. Zebiak and Cane,
1987) are still widely used, for example to look
at interactions between ENSO and other climatic
forcings such as insolation change (see Chapter 9,
section 9.5.3).
1.4 Drivers of tropical
environmental change
In the tropics, as elsewhere, the recognition of insolation as a driver of long term climatic change has
been of profound importance. The variations in
incoming radiation, which have come to be known
as Milankovitch cycles, are referred to in many
chapters, so a brief introduction to them and in
particular their importance in tropical latitudes is
given here. The theoretical framework for these
cycles came from James Croll in the 1860s, but has
come to be much more closely associated with
Milutin Milankovitch (1941). The cycles of eccentricity, tilt and precession (100 kyr, 41 kyr and
∼23 kyr respectively) affect both the distribution
17
and/or amount of incoming solar radiation (insolation). The net effect of these cycles on insolation at
20 °N and °S is illustrated in Fig. 1.6. The changing
amplitudes over the last 160 kyr and the clear evidence of the precession cycle are notable.
Milankovitch proposed that changes in insolation due to the 41 kyr cycle and the 23 kyr cycle
could drive NH ice growth (and hence explain ice
ages). Work on marine core isotope records starting
in the 1970s confirmed the presence of these periodicities (Hays et al., 1976). The ubiquity of the
cycles in records of all kinds is now well known,
but questions of how they are transmitted and
amplified over different timescales remain (see
Zachos et al., 2001 and Ruddiman, 2006, for overviews). In the context of this volume it is also
notable that long cores from the tropical oceans
played an important role in extending the marine
timescale based on orbitally-driven insolation inputs,
with the oxygen isotope (δ18O) signal (adjusted for
lags in response to insolation changes), forming the
basis for the SPECMAP timescale (Imbrie et al.,
1984; Shackleton et al., 1990).
The strongest direct effect on the tropics comes
from precession (see Fig. 1.7), which alters the
intensity of summer insolation and hence seasonality (although over a full year insolation anomalies
sum to zero). Stronger seasonality in the tropics,
associated with changes in monsoon strength, gives
an antiphase signal between the NH and Southern
Hemisphere (SH). The pattern of precession-driven
changes in the position of the ITCZ and in monsoon
strength is sometimes referred to as the ‘typical
tropics’. Clement et al. (2004) suggest that the precession cycle is vital to understanding climatic
change in the tropics, particularly the hydrological
response.
A dominance of precession (or half precession)
cyclicity is reported from a number of tropical sites,
including Lake Naivasha, Kenya (Trauth et al.,
2003), Sanbao Cave, China (Wang, Y et al., 2008),
Carnegie Ridge, Panama Basin (Pena et al., 2008)
and Botuvera Cave, Brazil (Wang, X et al., 2007),
although it should be noted that precession cycles
are not clear in Australian records (see Chapter 7,
section 7.8.2). Whether the coupling of monsoon
strength with precession holds through time has
18
(a)
Chapter 1
550
LGM
LIG
500
W/m2
450
400
350
300
250
0
(b)
20
40
60
80
kyr BP
100
120
140
160
140
Fig. 1.7 Strength of insolation to
the upper atmosphere (a) at 20 °N
and (b) at 20 °S over the last
160 160 kyr, showing the 19–23 kyr
influence of the precession cycle.
Data from Analyseries (Berger,
1992; Paillard et al., 1996).
550
500
W/m2
450
400
350
300
250
0
20
40
60
80
kyr BP
100
Summer
Winter
been questioned by Fritz et al. (2004) based on
results from the Salar de Uyuni (see Chapter 8,
section 8.4.3). Precession is also recorded by the
methane (CH4) in ice cores (e.g. Brook et al., 2000),
mainly because of the importance of tropical wetlands as a source. Changes in methane emissions
are seen as part of an early (quick) response to
120
insolation change. The role of the tropics in greenhouse gas emissions is discussed further below.
In addition to the link to the monsoon, changes
in precession are also associated with changes in
ENSO. The possibility that strong precession-driven
insolation during the early Holocene suppressed
El Niño has been mentioned above (Nederbragt
Introduction
and Thurow, 2005; Peterson and Haug, 2006).
A stronger monsoon was associated with a more
La Niña-like state. Modelling (Clement et al., 1999)
using the Zebiak–Cane model and precession
forcing showed an ENSO-like response to precession (but not obliquity). These authors also
suggest that modelled changes in ENSO would be
a mechanism for generating a globally synchronous
response to Milankovitch forcing (via changes in
Hadley cell strength and heat and moisture transport to high latitudes).
Although precession is an obvious focus for
driving millennial scale change in the tropics, it is
also linked to eccentricity and tilt. Precession is
modulated by eccentricity; low eccentricity reduces
the impact of precession on insolation. Overall,
orbital cycles produce a variability in tropical
insolation of around ±8% to the upper atmosphere,
but at times of higher eccentricity (e.g. at the
LGM), the precession signal would have been
stronger than it is under low eccentricity conditions (modern). Tilt, unlike precession, does change
annual insolation receipt at a given latitude and
is important for monsoons due to its impact of
warming over continental interiors. The importance of this is evident in the results of modelling
experiments (e.g. Prell and Kutzbach, 1992; see
Chapter 9, section 9.2.4 for further discussion).
Some tropical (especially marine) records are actually dominated by the 41 kyr (tilt) and 100 kyr
(eccentricity) signals (e.g. Lawrence et al., 2006;
Medina-Elizalde and Lea, 2010 from either side of
the Pacific), perhaps because the precession cycle
itself is too short to affect the oceanic heat budget
(Philander and Fedorov, 2003).
The influence of orbital variability as a pacemaker of tropical climate change was subject to
modification by other factors. Differences in cloudiness, the presence of large ice masses or changes
in ocean circulation could easily have exerted
strong local climate influences that disrupted the
orbital signature. Alternatively these influences
could have magnified the ecological effects of the
change in insolation. A last point to make about
these gradual changes in insolation is that they can
cause abrupt changes in ecosystems. Modelled and
19
empirical data suggest that even with a gradual
change in climate, a tipping point can be reached
that results in very fast and strong ecological change
(Scheffer et al., 2001, 2009; Bush et al., 2010).
Another form of solar variability that is expressed
in higher resolution tropical records is sunspot
cycles. Although the exact mechanisms through
which these small changes in solar output are propagated to drive global climate change are not clear
(low solar activity seems to be associated with a
more equatorward position of the ITCZ, weaker
Hadley cell and monsoon circulations, i.e. drier in
the tropics; van Loon et al., 2004; Versteegh, 2005),
these solar cycles are being reported with increasing frequency. A cycle of around 200 years (the
Suess Cycle) seems to be particularly common.
Examples of solar variability records in tropical latitudes include the work of Hodell et al. (2001) and
Stager et al. (2005).
Even as the magnitude of Quaternary change in
tropical latitudes became increasingly apparent,
there was still a view that the tropics were essentially ‘passive’ in the global climate system, simply
responding to changes driven by high latitudes and
perhaps particularly in and around the North
Atlantic. This view has only started to change fundamentally in the last 10–15 years as the importance of the tropical oceans and tropical controls
on greenhouse gas concentrations have become
increasingly evident (see Kerr, 2001). One of
reasons behind the persistence of this ‘passive’ view
may have been the highly influential CLIMAP SST
reconstructions (CLIMAP, 1976, 1981). These suggested little change in tropical ocean SSTs (about
1 °C) and as these values were used to prescribe
SSTs in early GCM experiments for the LGM they,
unsurprisingly, showed little effect. In contrast
tropical continental reconstructions (largely pollen,
but also noble gases) showed major temperature
depression at the LGM (5–8 °C). Guilderson et al.
(1994) reconstructed SSTs based on Barbados corals
and showed a temperature reduction of 5 °C at
19 kyr BP. Their data showed major and rapid
changes in tropical SSTs and it was suggested that
these, with changes in CO2 concentrations and in
the thermohaline circulation, could explain the
20
Chapter 1
apparent synchroneity of global change in the
period after the LGM. Reconstructing LGM tropical
SSTs at the global scale has been problematic. One
of the difficulties seems to be that there are no
modern analogues for some LGM foraminiferal
assemblages that form the basis for the temperature
reconstructions. In a reanalysis of the CLIMAP
data, Mix et al. (1999) addressed this problem and
determined that the original analysis had underestimated SST change, by significant amounts in
some places. Their new analysis suggested that the
East Pacific cold tongue had actually cooled by 6 °C
at the LGM. Concern about sources of error in
faunal based tropical and subtropical SSTs leading
to bias led to the development of bias-adjusted SST
estimates and exploration of these on modelled
climate (e.g. Hostetler et al., 2006). Other more
recent studies of SST change include that of
Kucera et al. (2005) based on a multiproxy approach
and showing significant tropical ocean cooling, and
that of Ballantyne et al. (2005) who undertook
a Bayesian meta-analysis and identified tropical
SST cooling of 2.7 ± 0.5 °C. See Chapter 3, sections
3.2.1 and 3.4.1 for further discussion of SST
reconstructions.
Given the known significance of ENSO, driven
by tropical Pacific SSTs, in causing climate change
at the global scale, it is perhaps surprising that
recognition of the tropical oceans as drivers of
change, over a range of timescales, has been slow
to develop. Cane (1998) suggested that the tropical
Pacific could drive both orbital and millennial
scale climate variability, using ENSO teleconnections as an example. Long term ENSO type variability could favour interglacial/interstadials (El Niño/
warm phase) or glacial/stadials (cold phase). Even
recent changes in the North Atlantic/European
climate have been attributed to changes in tropical
SSTs (specifically in the Indo-Pacific region) (Hoerling et al., 2001). Broecker (2003) considered the
relative roles of the Atlantic thermohaline circulation and changes in the tropical atmosphere–ocean
system (basically ENSO) in driving abrupt climate
change, focusing on D–O events and the timing of
warming, and Heinrich (H) Events. The Younger
Dryas (sometimes called H0) was seen as a distinct
event, almost certainly driven by meltwater dis-
charge into the North Atlantic due to ice sheet
dynamics, but Broecker recognised that warming
prior to D–O events might suggest a tropical trigger.
These early warming signals prior to H Events have
been detected in sites far removed from the North
Atlantic such as the Santa Barbara Basin off California (Hendy et al., 2002).
1.5 The tropics as drivers
of change
Although the focus above has been largely on
factors that might drive climate change in the
tropics, there are a number of climate forcings that
originate in, or are significantly controlled by, the
tropics. In simple energy terms, the tropical atmosphere and oceans are a major energy source, as
they help to redistribute incoming solar radiation
(see Chapters 2 and 3). In addition, the tropics are
important in controlling a number of long lived
radiatively active gases and aerosols, including
water vapour. Here the role of the tropics in regulating the greenhouse gases carbon dioxide (CO2),
methane (CH4) and nitrous oxide (N2O) (positive
forcings), and the impact of tropical volcanoes in
sulphate aerosol production and deserts as mineral
dust sources (negative forcings), are summarised.
1.5.1 The tropics and greenhouse gas
concentrations
The tropics play a major role in terms of both
sources and sinks of CO2. The major tropical CO2
sources are land use change and deforestation; over
the last 20 years the IPCC (2007) estimate that the
CO2 flux due to land use change (1.6 GtC yr-1) has
been dominated by tropical deforestation. Although
this may be balanced by uptake by tropical live
biomass, it is not well quantified. Annual variability
in CO2 fluxes, such as that evident in the Mauna
Loa (Hawaii) record, is strongly influenced by tropical land areas. The importance of tropical areas
(specifically the forests) as CO2 sinks and the potentially dramatic impacts of future warming (and
drying) have been explored using GCMs (e.g. Cox
et al., 2000). Friedlingstein et al. (2006) provide a
more recent estimate of the possible feedbacks. The
Introduction
tropical oceans also play a part and are thought to
outgas CO2 to atmosphere (estimated mean flux
∼0.7 GtC yr-1). Upwelling waters are particularly
CO2 rich and the highest CO2 emissions seem to
occur in the equatorial central and eastern
Pacific. Both terrestrial and oceanic CO2 fluxes are
affected by El Niño. On land, sources increase
(drier, more forest fires), but oceanic sources
decrease as upwelling in the eastern Pacific is
reduced (see above).
Biogenic sources dominate CH4 emissions (> 70%)
and natural CH4 comes mainly from wetlands,
both tropical and boreal, where emissions are
associated with the occurrence of anaerobic
conditions. According to Loulergue et al. (2008)
the contribution is about one third boreal and
two thirds tropical, and they suggest that in the
pre-industrial period wetlands may have provided
85% of the total source. Other major biogenic
sources include rice agriculture (paddies), biomass
burning and ruminants; again these have significant
contributions from tropical areas. δ13C can be used
to fingerprint sources of CH4 and this has been
used extensively in relation to ice core records
(see below). Keppler et al. (2006) have suggested
that it may also be possible for CH4 to be emitted
by living vegetation (primarily from tropical
forests and grasslands) under aerobic conditions.
CH4 emissions are particularly sensitive to
climate change via temperature and precipitation
(moisture). Methane sinks are dominated by OH
and the more water vapour there is, the more OH
is produced. The abundance of OH is also affected
by large volcanic eruptions (see below), which
subsequently affects CH4 concentrations.
Uncertainties in estimates of N2O sources are
even larger than for CO2 and CH4, but tropical soils
are important sources, probably contributing more
than 50% of N2O emissions (Prinn et al., 1990;
Hirsch et al., 2006). Other sources include agriculture (especially with fertiliser use) and the oceans
(through denitrification). Hirsch et al. (2006)
estimate that about 26% of N2O comes from the
oceans.
The records of greenhouse gas concentrations in
ice cores have played a large part in our developing
understanding of the association between atmos-
21
pheric composition and climate. Particularly in the
case of CH4, it is clear that the tropics are a major
methane source. The high resolution Greenland
(GRIP) and Antarctic (Byrd) ice core CH4 records
(Chappellaz et al., 1997) have been interpreted in
terms of changes in source strengths in tropical
and boreal wetlands. Up to about 5 kyr BP, the
dominant source seems to have been the tropical
wetlands, with particularly high values between
11.5 and 9.5 kyr BP. Drier conditions in the tropics
between about 7 and 5 kyr BP were apparently
reflected in lower overall CH4 concentrations. Dallenbach et al. (2000) extended the study into the
last glacial (to 46 kyr BP) and demonstrated that
mean CH4 levels in the GRIP core at the LGM were
362 ppbv compared with an average of 654 ppbv
over the available Holocene record. These studies
also explore interpolar differences, which were at
a minimum at the LGM and a maximum during
warm periods, especially 5–2.5 kyr BP. Dallenbach
et al. (2000) attribute the larger portion of total
CH4 during much of the last glacial to northern
high latitudes, except at the LGM, but tropical
sources predominate during cold periods (although
totals are low). Wolff and Spahni (2007) revisited
the ice core CH4 and N2O records and place more
emphasis on changes in sinks which they see as
necessary to explain the magnitude of observed
changes in concentration. The authors also used
isotopic measurements (δ13C) to help discriminate
between sources; biogenic sources are isotopically
light, whilst fossil fuels and biomass burning produces an isotopically heavier signal. Current CH4
levels exceed any recorded over the last 650 000
years.
The Holocene story of CH4 is complex. Ruddiman
(2003) suggested that human impacts on global
CH4 could have begun as early as 5 kyr BP, largely
due to activity in the tropics (rice paddies etc.),
although his interpretation of the records has been
disputed (Sowers, 2009; Singarayer et al., 2011).
Interestingly, Singarayer et al. (2011) conclude that
the main driver of the observed late Holocene
increase in CH4 is emissions from the Southern
Hemisphere tropics, driven by precession. N2O
records are not as complete as CH4; they show a
similar overall trend, but recent change is not as
22
Chapter 1
extreme. Some source discrimination is possible
using δ15N (agricultural sources, especially fertilisers, are isotopically light).
The link between tropical climate and methane
and N2O emissions is illustrated by the correlation
of CH4 with precession cycles (and monsoons); this
correlation persists on millennial timescales. Ivanochko et al. (2005) suggest that changes in the tropical hydrological cycle during stadials/interstadials
(D–O warm interstadials and cold stadials) could
drive temperature changes in northern wetlands
and hence changes in CH4 emissions. Changes in
monsoon strength would also drive changes in N2O
emissions from the oceans (via denitrification) and
from terrestrial wetlands. δ15N values in marine
sediments can serve as a proxy for denitrification
and loss of N2 and N2O to the atmosphere. They see
impacts of changes in these greenhouse gas emissions as a tropical mechanism for amplifying and
perpetuating millennial scale climatic changes initiated in the North Atlantic (D–O cycles).
Loulergue et al. (2008) describe an 820 kyr record
from EPICA Dome C. They propose that tropical
sources and sinks are the dominant control of
the CH4 budget (linked to changes in monsoon
strength), and that boreal wetlands are only significant during major terminations. Their CH4 record
shows a dominant 100 kyr (eccentricity) cycle until
420 kyr BP, then a stronger precession signal.
1.5.2 Impacts of low latitude volcanic
eruptions
While the tropics clearly help to control warming
greenhouse gas concentrations, eruptions of tropical volcanoes that emit large quantities of SO2
and SO4 (leading to the formation of sulphate
aerosol) are most likely to have global and persistent cooling impacts on climate. Aerosols reduce
radiative forcing (dimming) and are associated with
general cooling, although warming may occur over
NH continents in the first winter after an eruption.
The lifetime of aerosols is quite short, so the effects
based on observations only last about 2 years.
Other impacts of these large eruptions (including
those arising from emissions of water vapour,
vegetation/albedo effects and other alterations to
atmospheric chemistry), are more complicated.
Robock (2000) provides a general overview of the
effects of eruptions on atmospheric inputs and
radiative forcing.
It is clear that the eruption of large tropical volcanoes provides a possible mechanism to explain
abrupt climatic change. The most recent example
is Mt Pinatubo (Philippines) which erupted in
1991. This had a VEI (Volcanic Explosivity Index)
of 6, and emitted about 20 Mt SO2 into the stratosphere (Robock, 2002). The duration of any climatic impacts arising from really large eruptions is
still unclear. The super-eruption of Toba (northern
Sumatra) about 73 kyr BP, with a VEI of 8, may
have been the largest volcanic event of the last
2 Myr (certainly of the last 100 kyr), and its effects
were seen globally (e.g. in GRIP). The duration of
its impact is disputed, but Williams et al. (2009),
based on pollen and isotopic data, suggest that it
caused significant cooling and drying over 2 kyr
and had a severe global impact upon the biosphere,
hydrosphere and humans. The record of this
massive eruption is discussed further in Chapter 6,
section 6.3.3.
D’Arrigo et al. (2009) have explored the impact
of volcanic forcing on temperatures within the
tropics over the last 400 years. They developed
a zonally averaged annual tropical temperature
record for the tropics between 30 °N and °S, using
a range of proxies, and compared this to an index
of volcanic forcing. They found that tropical temperatures were affected more by tropical eruptions
than by those occurring at high latitudes (even
large ones), although the magnitude of the response
was smaller than at high latitudes. They noted that
the most sustained cool period in their record
occurred in the early nineteenth century (especially 1815–1818), which included the eruption of
Tambora in 1815 (although they acknowledge that
this was also a period of low solar irradiance). In
contrast, Oman et al. (2006) found evidence of the
impact of high latitude eruptions in the tropics,
specifically on Nile River flow via changes in the
monsoon following the eruption of Laki (Iceland)
in 1783–1784. It is, however, interesting to note
Introduction
that although 1782–1783 was a strong El Niño
period (see Ortlieb, 2000) this is not considered by
Oman et al. In contrast, in their study, D’Arrigo
et al. (2009) tried to filter out the impact of ENSO,
but showed little effect on composite temperature
(which masks spatial impacts).
Another area of interest has been the possible
ENSO type response to volcanic eruptions, originally suggested by Handler (1984) (see Adams
et al., 2003). This was followed up by Emile-Geay
et al. (2008) who modelled the impact of tropical
eruptions on ENSO using the Zebiak–Cane model
(see above), with a particular focus on a very large
eruption of about 1258. They suggested that eruptions larger than that of Mt Pinatubo could increase
the likelihood of an El Niño occurring (but not its
intensity) whilst recognising that El Niños can be
forecast using oceanic/atmospheric data alone.
They proposed that the 1258 eruption may have
caused a moderate/strong El Niño in the midst of
a period dominated by La Niña-like conditions (the
Mediaeval Climatic Anomaly). It is worth noting,
however, that Robock (2000) does not support a
causal link between eruptions and ENSO.
One further aspect of volcanic eruptions is their
effect on stratospheric ozone depletion. It is evident
that recent large eruptions (such as Pinatubo)
have impacted upon stratospheric O3 levels, since
sulphate aerosol inputs to the stratosphere react
with chlorine (largely anthropogenic in origin) to
provide sites for catalytic reactions that lead to
ozone depletion. Robock (2002) suggests that, as
this process relies on the presence of chlorine, it
would not have occurred prior to the mid twentieth century.
1.5.3 Dust emissions from the tropics
and subtropics
In recent years, the correlation between glacial
periods and the flux of mineral dust to the
atmosphere has attracted considerable interest.
Model and data compilations (e.g. from marine
sediments, polar and tropical ice caps, and
continental loess deposits) indicate that much of
the world experienced increased dust deposition
23
around the LGM, with global average mineral dust
loading in the atmosphere around 50% higher
than in preindustrial times (Kohfeld and Harrison,
2001). However, some areas, most notably the
tropics and the poles, experienced higher loadings.
For example, dust fluxes from Africa to the tropical
and subtropical Atlantic during the LGM were 3–5
times higher than modern values (see Chapter 4,
section 4.4), whilst fluxes into the North Pacific
from the Americas and east Asia were 1–2 times
higher (see Harrison et al., 2001). Modelling studies
suggest that such elevated dust levels may have
induced an average cooling of up to 0.72 °C in
surface air temperature over the tropical oceans
(Yue et al., 2011).
Records from the Antarctic Vostok and Dome C
ice cores show even more dramatic increases in dust
deposition during glacials, with 10–12 times larger
fluxes of dust at glacial maxima relative to the mean
flux and 27–30 times larger fluxes of dust during the
LGM compared to the present day (Petit et al., 1999;
Delmonte et al., 2004). Analyses of mineralogical
and isotopic tracers have been used to suggest that
glacial dust in these cores is likely to have been
transported from Patagonia (Grousset et al., 1992;
Basile et al., 1997) with an Australian contribution
during interglacials (e.g. Delmonte et al., 2004, 2008).
The degree to which atmospheric dust loading is
a response to, or a contributory cause of, climate
changes on glacial–interglacial timescales is still
uncertain (Harrison et al., 2001; Bar-Or et al.,
2008). Glaciation, for example, has been suggested
to increase dust flux into the atmosphere by (i)
enhancing the effects of continentality (due to
lower sea level), (ii) increasing the potential for soil
erosion (through a reduction of vegetation cover
caused by lower moisture availability) and (iii)
increasing wind speeds (due to steeper pole–
Equator pressure gradients) (e.g. Harrison et al.,
2001; Harrison and Prentice, 2003). However,
increased dust flux can also have both positive
and negative feedbacks on glaciations through the
aerosol direct radiative effect (see Yoon et al.,
2005), the effect of dust deposition on snow and
ice albedo, and the impact of aerosol particles on
the reflection and absorption properties of clouds
24
Chapter 1
(Rosenfeld et al., 2006; Bar-Or et al., 2011). Dust
transported to the oceans can also affect climate
indirectly by modulating the supply of elements
such as bioavailable iron, a micronutrient essential
to photosynthesis in phytoplankton (Martin et al.,
1991). Through this mechanism, variations in dust
flux can influence the uptake of carbon in marine
ecosystems and, in turn, the atmospheric concentration of CO2 (Maher et al., 2010).
The relationship between changes in atmospheric dust loading and other palaeoenvironmental
indicators from ice cores is not straightforward. For
example, CO2 concentrations had already reached
near-glacial levels by the time dust concentrations
in the Vostok ice core began to increase around
65 kyr BP (Petit et al., 1999). In contrast, the
decrease in dust loading evident in the Vostok core
appears to be synchronous with, or even to precede,
the increases in atmospheric CO2 concentrations
during deglaciations (Harrison et al., 2001).
1.6 Extra-tropical forcing
As described in a number of chapters in this volume,
environmental changes affecting the North Atlantic
(e.g. D–O cycles and H Events, the YD, changes
in thermohaline circulation) remain the primary
focus of research into the origins of millennial scale
climate variability. Such changes are recognised
well beyond the North Atlantic and are present in
many tropical records of sufficient resolution (see
examples in Chapters 4, 6, 7 and 8). Initially, interest centred upon the Younger Dryas (c. 13 kyr or
12.6–11.5 kyr BP). Evidence for the YD and earlier
H Events (marking the end of D–O cycles) is widespread in the tropics and subtropics (e.g. Baker
et al., 2001; Lea et al., 2003; Peterson and Haug,
2006; Wang X et al., 2006, 2007; Wang, Y et al.,
2008). The near synchroneity of these events with
the chronology of the Greenland ice cores has led
to the suggestion that the tropics played an active
role in propagating signals from the North Atlantic,
probably via changes in the location of the ITCZ,
monsoon strength and methane emissions (see
above). The expression of these events (e.g. wetter/
drier) is, however, spatially variable. For example,
at Botuvera in Brazil, the YD was wetter than
present, but at Hulu in China it was drier (Wang,
Y et al., 2001). This serves to highlight the complexity of response to climate forcings and the need
to consider specific locations within the climate
system. It is a timely reminder of the pitfalls of
assuming that patterns are simply replicated, as was
the case when the early evidence of lake level rise
in the southwest USA during glacials was assumed
to apply to all lower latitude locations (see section
1.2.1 above).
Changes in the high latitudes of the Southern
Hemisphere (primarily the Southern Ocean) have
also been considered drivers of millennial scale
climate variability, particularly in relation to
warming during the early phases of deglaciation
(terminations) caused by insolation forcing from
the Southern Hemisphere and its effects on the
global CO2 budget (e.g. Broecker and Henderson,
1998; Shulmeister et al., 2006). Many of the
regional chapters (e.g. Chapters 4, 7 and 8) include
examples of tropical records which appear to show
changes more consistent with those over Antarctica
than the North Atlantic. Although most of these
are from the southern tropics, there are a number
of NH sites which also seem to show a SH influence
(e.g. Williams et al., 2010).
1.7 Organisation of the volume
Quaternary Environmental Change in the Tropics is
organised into three sections. Section A (‘Global
contexts’) includes this introduction plus an overview of the contemporary climatology of the tropics
(Chapter 2: Stefan Hastenrath). The latter chapter
is designed to provide a background to the major
features of tropical climate zones, with specific features of regional climate developed more fully in
each of Chapters 3 to 8.
Section B (‘Regional environmental change’)
contains six substantive chapters. These review
the evidence for environmental changes in the
tropical oceans (Chapter 3: Jan-Berend W. Stuut,
Matthias Prange, Ute Merkel and Silke Steph),
Africa (Chapter 4: David Nash and Mike Meadows),
India, Arabia and adjacent areas (Chapter 5: Ashok
Introduction
Singhvi, Nilesh Bhatt, Ken Glennie and Pradeep
Srivastava), China and Southeast Asia (Chapter 6:
Dan Penny), Australia and the southwest Pacific
(Chapter 7: Peter Kershaw and Sander van der
Kaars) and Latin America and the Caribbean
(Chapter 8: Mark Bush and Sarah Metcalfe).
The authors of each of the regional chapters
were requested to address a series of specific
issues within their reviews. First, they were asked
to summarise the evidence for environmental
change in their specific region, making reference
to available sedimentological, geochemical (including isotopic), biological, geomorphological and
archaeological evidence spanning the entire Quaternary. Second, they were asked to highlight
any issues of spatial (e.g. longitudinal, Northern
vs. Southern Hemisphere) and temporal (e.g.
glacial vs. interglacial) variability within and
between records for their region. Third, they were
requested to consider the drivers of environmental
changes, drawing attention to climatic versus
human-induced forcing mechanisms where appropriate. We consider that all have more than adequately met this brief.
The volume concludes with Section C (‘Global
syntheses’) which contains three chapters designed
to span the tropics and give a global perspective
on key issues. Chapter 9 (Zhengyu Liu and Pascale
Braconnot) reviews the contributions made by
modelling studies to our understanding of tropical
environments during the Quaternary. Chapter 10
(Georgina Endfield and Robert Marks) considers
the evidence for environmental change in the
tropics over the last 1000 years. Finally, Chapter
11 (David Nash and Sarah Metcalfe) draws
together the evidence for Quaternary environmental changes in the tropics in a global synthesis,
considers the impact of future climate change
upon tropical regions and identifies a number of
recommendations for areas of future research.
Acknowledgements
The authors would like to thank Elaine Watts
(School of Geography, University of Nottingham)
for producing the figures for this chapter.
25
References
Adams, J., Mann, M. and Ammann, C. (2003) Proxy
evidence for an El Niño like response to volcanic
forcing. Nature 426, 274–278.
Agassiz, L. (1840) Etudes sur les glaciers. Jent et Gassmann,
Neuchâtel, Switzerland.
Agassiz, L. (1868) A Journey in Brazil. Ticknor and Fields,
Boston, USA.
Aharon, P. and Chappell, J. (1986) Oxygen isotopes, sea
level changes and temperature history of a coral reef
environment in New Guinea over the last 105 years.
Palaeogeography, Palaeoclimatology, Palaeoecology 56,
337–379.
An, Z. (2000) The history and variability of the East Asian
paleomonsoon climate. Quaternary Science Reviews 19,
171–187.
Baillie, J.E.M., Hilton-Taylor, C. and Stuart, S.N. (2004)
2004 IUCN Red List of Threatened Species. A Global Species
Assessment. IUCN, Gland, Switzerland.
Baker, P.A., Rigsby, C.A., Seltzer, G.O. et al. (2001)
Tropical climate changes at millennial and orbital
timescales on the Bolivian Altiplano. Nature 409,
698–701.
Ballantyne, A.P., Lavine, M., Crowley, T.J., et al. (2005)
Meta-analysis of tropical surface temperatures during
the last glacial maximum. Geophysical Research Letters 32,
L05712.
Bar-Or, R., Erlick, C. and Gildor, H. (2008) The role of
dust in glacial-interglacial cycles. Quaternary Science
Reviews 27, 201–208.
Bar-Or, R., Gildor, H. and Erlick, C. (2011) The aerosolBenard cell effect on marine stratocumulus clouds and
its contribution to glacial-interglacial cycles. Journal of
Geophysical Research-Atmospheres 116, article number:
D10119, doi: 10.1029/2010JD014470.
Bard, E., Hamelin, B. and Fairbanks, R.G. (1990) U-Th
ages obtained by mass spectrometry in corals from Barbados: sea level during the past 130,000 years. Nature
346, 456–458.
Basile, I., Grousset, F.E., Revel, M., et al. (1997) Patagonian origin of glacial dust deposited in East Antarctica
(Vostok and Dome C) during glacial stages 2, 4 and 6.
Earth and Planetary Science Letters 146, 573–589.
Beerling, D.J. and Woodward, F.I. (1993) Ecophysiological responses of plants to global environmental change
since the Last Glacial Maximum. New Phytologist 125,
641–648.
Berger, A. (1992) Astronomical theory of paleoclimates
and the last glacial-interglacial cycle. Quaternary Science
Reviews 11, 571–581.
26
Chapter 1
Betancourt, J.L., Van Devender, T.R. and Martin, P.S.
(1990) Packrat Middens. The Last 40,000 Years of Biotic
Change. The University of Arizona Press, Tucson.
Birks, H.J. (2005) Quantitative palaeoenvironmental
reconstructions from Holocene biological data. In:
Mackay, A., Battarbee, R., Birks, J. and Oldfield, F. (eds)
Global Change in the Holocene. Hodder Arnold, London,
pp. 107–123.
Birks, H.J.B. and Birks, H.H. (1980) Quaternary Palaeoecology. Edward Arnold, London.
Braconnot, P., Otto-Bliesner, B., Harrison, S., et al. (2007)
PMIP2 coupled simulations of the mid-Holocene and
Last Glacial Maximum – part 1: experiments and large
scale features. Climate of the Past 3, 261–277.
Bridgman, H.A. and Oliver, J.E. (2006) The Global Climate
System, Patterns, Processes and Teleconnections. Cambridge
University Press, Cambridge.
Broecker, W.S. (2003) Does the trigger for abrupt climate
change reside in the ocean or in the atmosphere?
Science 300, 1519–1522.
Broecker, W.S. and Henderson, G.M. (1998) The sequence
of events surrounding Termination II and their implications for the cause of glacial-interglacial CO2 changes.
Paleoceanography 13, 352–364.
Brook, E.J., Harder, S., Severinghaus, J., et al. (2000) On
the origin and timing of rapid changes in atmospheric
methane during the last glacial period. Global Biogeochemical Cycles 14, 559–572.
Bush, M.B., Hanselman, J.A. and Gosling, W.D. (2010)
Nonlinear climate change and Andean feedbacks: an
imminent turning point? Global Change Biology 16,
3223–3232.
Candy I., Black S. and Sellwood B.W. (2004) Quantifying
timescales of pedogenic calcrete formation using Useries disequilibria. Sedimentary Geology 170, 177–187.
Cane, M.A. (1998) A role for the tropical Pacific. Science
282, 59–62.
Cane, M.A., Braconnot, P., Clement, A., et al. (2006)
Progress in paleoclimate modeling. Journal of Climate 19,
5031–5057.
Chappellaz, J., Barnola, J.-M., Raynaud, D., et al. (1990)
Atmospheric CH4 record over the last climatic cycle
revealed by the Vostok ice core. Nature 345, 127–131.
Chappellaz, J., Blunier, T., Kints, S., et al. (1997) Changes
in the atmospheric CH4 gradient between Greenland
and Antarctica during the Holocene. Journal of Geophysical Research 102, 15987–15997.
Chase, B.M., Meadows, M.E., Scott, L., et al. (2009) A
record of rapid Holocene climate change preserved in
hyrax middens from southwestern Africa. Geology 37,
703–706.
Clement, A.C., Seager, R. and Cane, M.A. (1999) Orbital
controls on the El Nino/Southern Oscillation and the
tropical climate. Paleoceanography 14, 441–456.
Clement, A.C., Cane, M.A. and Seager, R. (2001) An
orbitally driven tropical source for abrupt climate
change. Journal of Climate 14, 2369–2375.
Clement, A.C., Hall, A. and Broccoli, A.J. (2004)
The importance of precessional signals in the tropical
climate. Climate Dynamics 22, 327–341.
Clift, P.D. and Plumb, R.A. (2008) The Asian Monsoon.
Causes, History and Effects. Cambridge University Press,
Cambridge.
CLIMAP Project members (1976) The surface of ice-age
earth. Science 191, 1131–1137.
CLIMAP Project members (1981) Seasonal reconstructions of the earth’s surface at the last glacial maximum.
Geological Society of America Map and Chart Series
MC-36.
COHMAP Members (1988) Climatic changes of the last
18,000 years. Observations and model simulations.
Science 241, 1043–1052.
Cox, P.M., Betts, R.A., Jones, C.D., et al. (2000) Acceleration of global warming due to carbon cycle feedbacks
in a coupled climate model. Nature 408, doi:10.1038/
35041539.
Crowley, T.J., Kim, K-Y., Mengel, J.G. and Short, D.A.
(1992) Modeling 100,000 year climate fluctuations in
pre-Pleistocene time series. Science 225, 705–707.
D’Arrigo, R., Wilson, R. and Tudhope, A. (2009) The
impact of volcanic forcing on tropical temperatures
during the past four centuries. Nature Geoscience 2,
51–56.
Dallenbach, A., Blunier, T., Fluckiger, J., et al. (2000)
Changes in the atmospheric CH4 gradient between
Greenland and Antarctica during the Last Glacial and
transition to the Holocene. Geophysical Research Letters
27, 1005–1008.
Dallmeyer, A., Claussen, M. and Otto, J. (2010) Contribution of oceanic and vegetation feedbacks to Holocene
climate change in monsoonal Asia. Climate of the Past 6,
195–218.
Damuth, J.E. and Fairbridge, R.W. (1970) Arkosic sands
of the last glacial stage in the tropical Atlantic off Brazil.
Geological Society of America, Bulletin 81, 189–206.
de Beer, G. R. (1953) Bernard Friedrich Kuhn’s investigations on glaciers. Annals of Science 9, 323–341.
de Menocal, P.B. (1995) Plio-Pleistocene African climate.
Science 270, 53–59.
de Menocal, P.B. (2004) African climate change and
faunal evolution during the Pliocene-Pleistocene. Earth
and Planetary Science Letters 220, 3–24.
Introduction
de Saussure, H.-B. (1779) Voyages dans les Alpes,
précèdes d’un essai sur l’histoire naturelle des environs de
Genève, Vol. 1. Barde, Manget and Compagnie, Geneva,
Switzerland.
Delmonte, B., Basile-Doelsch, I., Petit, J.-P., et al. (2004)
Comparing the EPICA and Vostok dust records during
the last 220,000 years: stratigraphical correlation and
provenance in glacial periods. Earth-Science Reviews 66,
63–87.
Delmonte, B., Andersson, P.S., Hansson, M., et al. (2008)
Aeolian dust in East Antarctica (EPICA-Dome C and
Vostok): provenance during glacial ages over the last
800 kyr. Geophysical Research Letters 35, L07703.
doi:10.1029/2008GL033382.
Diaz, H. and Markgraf, V. (eds) (1992) El Niño, Historical
and Paleoclimatic Aspects of the Southern Oscillation. Cambridge University Press, Cambridge.
Diaz, H. and Markgraf, V. (eds) (2000) El Niño and
the Southern Oscillation, Multiscale Variability and Global
and Regional Impacts. Cambridge University Press,
Cambridge.
Diester-Haas, L. (1976) Late Quaternary variations in
northwest Africa deduced from east Atlantic sediment
cores. Quaternary Research 6, 299–314.
Drake, N.A., Blench, R.M., Armitage, S.J., et al. (2011)
Ancient watercourses and biogeography of the Sahara
explain the peopling of the desert. Proceedings of the
National Academy of Science 108, 458–462.
Emile-Geay, J., Seager, R., Cane, M.A., et al. (2008) Volcanoes and ENSO over the past millennium. Journal of
Climate 21, 3134–3148.
Emiliani, C. (1955) Pleistocene temperatures. Journal of
Geology 63, 538–575.
Fairbanks, R.G. (1989) A 17,000-year glacio-eustatic sea
level record: influence of glacial melting rates on the
Younger Dryas event and deep-ocean circulation.
Nature 342, 637–642.
Fedorov, A.V., Dekens, P.S., McCarthy, M., et al. (2006)
The Pliocene paradox (mechanisms for a permanent El
Niño). Science 312, 1485–1489.
Flenley, J.R. (1979) The Equatorial Rain Forest: A Geological
History. Butterworths, London.
Friedlingstein, P., Cox, P., Betts, R., et al. (2006) Climatecarbon cycle feedback analysis: results from the
C4MIP model intercomparison. Journal of Climate 19,
3337–3353.
Fritz, S.C., Cumming, B.F., Gasse, F. and Laird, K.R. (1999)
Diatoms as indicators of hydrologic and climatic change in
saline lakes. In: Stoermer, E.F. and Smol, J.P. (eds) The
Diatoms: Applications for the Environmental and Earth Sciences. Cambridge University Press, Cambridge, pp. 41–72.
27
Fritz, S.C., Baker, P.A., Lowenstein, T.K., et al. (2004)
Hydrologic variation during the last 170,000 years in
the southern hemisphere tropics of South America.
Quaternary Research 61, 95–104.
Fujioka, T., Chappell, J., Honda, M., et al. (2005) Global
cooling initiated stony deserts in central Australia 2–4
Ma, dated by cosmogenic 21Ne-10Be. Geology 33,
993–996.
Garnier, B.J. (1958) Some comments on defining the humid
tropics. Research Note 11, Department of Geography,
University of Ibadan.
Gates, W.L. (1976a) Modelling the ice-age climate. Science
191, 1138–1144.
Gates, W.L. (1976b) The numerical simulation of ice-age
climate with a global general circulation model. Journal
of Atmospheric Sciences 33, 1844–1873.
Gergis, J.L. and Fowler, A.M. (2009) A history of ENSO
events since A.D. 1525: implications for future climate
change. Climatic Change 92, 343–387.
Gibbard, P., Head, M.J. and Walker, M.J.C. (2010) Formal
ratification of the Quaternary System/Period and the
Pleistocene Series/Epoch with a base at 2.58 Ma. Journal
of Quaternary Science 25, 96–102.
Gilbert, G.K. (1890) Lake Bonneville. US Geological Survey
Monograph 1. USGS.
Goudie, A.S. (1972) The concept of post-glacial progressive desiccation. Research Papers, School of Geography, University of Oxford, pp. 4–48.
Goudie, A.S. (1999) The Ice Age in the tropics and its
human implications. In: Slack, P.A. (ed.) Environments
and Historical Change. Oxford University Press, Oxford,
pp. 10–32.
Goudie, A.S., Allchin, B. and Hegde, K.T.M. (1973) The
former extensions of the Great Indian Sand Desert.
Geographical Journal 139, 243–257.
Grab, S.W. and Nash, D.J. (2010) Documentary evidence
of climate variability during cold seasons in Lesotho,
southern Africa, 1833–1900. Climate Dynamics 34,
473–499.
Groot, M.H.M., Bogota, R.G., Lourens, L.J., et al. (2011).
Ultra-high resolution pollen record from the northern
Andes reveals rapid shifts in montane climates within
the last two glacial cycles. Climate of the Past 7,
299–316.
Grousset, F.S., Biscaye, P.E., Revel, M., et al. (1992) Antarctic (Dome C) ice-core dust at 18 k.y. B.P.: isotopic
constraints on origins. Earth and Planetary Science Letters
111, 175–182.
Grove, A.T. (1958) The ancient erg of Hausaland, and
similar formations on the south side of the Sahara.
Geographical Journal 124, 528–533.
28
Chapter 1
Grove, A.T. (1969) Landforms and climate change in the
Kalahari and Ngamiland. Geographical Journal 135,
192–212.
Grove, A.T. and Goudie, A.S. (1971) Late Quaternary lake
levels in the Rift Valley of southern Ethiopia and elsewhere in tropical Africa. Nature 234, 403–405.
Grove, A.T. and Warren, A. (1968) Quaternary landforms
and climate on the south side of the Sahara. Geographical Journal 134, 194–208.
Grove, A.T., Street, F.A. and Goudie, A.S. (1975) Former
lake levels and climatic change in the rift valley of
southern Ethiopia. Geographical Journal 141, 177–194.
Guilderson, T.P., Fairbanks, R.G. and Rubenstone, J.L.
(1994) Tropical temperature variations since 20,000
years ago: modulating interhemispheric climate change.
Science 263, 663–665.
Hahn, D.G. and Manabe, S. (1975) The role of mountains
in the south Asian monsoon circulation. Journal of
Atmospheric Science 32, 151–154.
Handler, P. (1984) Possible association of stratospheric
aerosols and El Nino type events. Geophysical Research
Letters 11, 1121–1124.
Harris, N.B.W. (2006) The elevation history of the Tibetan
Plateau and its implications for the Asian monsoon.
Palaeogeography, Palaeoclimatology, Palaeoecology 241,
4–15.
Harrison, S.P. and Prentice, I.C. (2003) Climate and CO2
controls on global vegetation distribution at the last
glacial maximum: analysis based on palaeovegetation
data, biome modelling and palaeoclimate simulations.
Global Change Biology 9, 983–1004.
Harrison, S.P., Kohfeld, K.E., Roelandt, C. and Claquin, T.
(2001) The role of dust in climate changes today, at the
last glacial maximum and in the future. Earth Science
Reviews 54, 43–80.
Hays, J.D., Imbrie, J. and Shackleton, N.J. (1976) Variations in the earth’s orbit: pacemaker of the ice ages.
Science 194, 1121–1132.
Hendy, I.L., Kennett, J.P., Roark, E.B. and Ingram, B.L.
(2002) Apparent synchroneity of submillennial scale
climate events between Greenland and Santa Barbara
basin, California from 30 – 10 ka. Quaternary Science
Reviews 21, 1167–1184.
Hirsch, A.I., Michalak, A.M., Bruhwiler, L.M., et al.
(2006) Inverse modeling estimates of the global nitrous
oxide surface flux from 1998–2001. Global Biogeochemical Cycles 20, GB1008, doi:10.1029/2004GB
002443.
Hodell, D.A., Brenner, M, Curtis, J.H. and Guilderson, T.
(2001) Solar forcing of drought in the Maya lowlands.
Science 292, 1367–1370.
Hoerling, M.P., Hurrell, J.W. and Xu, T. (2001) Tropical
origins for recent North Atlantic climate change. Science
292, 90–92.
Holmgren, C.A., Norris, J. and Betancourt, J.L. (2007)
Inferences about winter temperatures and summer
rains from the late Quaternary record of C-4 perennial
grasses and C-3 desert shrubs in the northern Chihuahuan Desert. Journal of Quaternary Science 22,
141–161.
Holmes, J.A. and Engstrom, D.R. (2005) Non-marine
ostracod records of Holocene environmental change.
In: Mackay, A., Battarbee, R., Birks, J. and Oldfield, F.
(eds) Global Change in the Holocene. Hodder Arnold,
London, pp. 310–327.
Hostetler, S., Giorgi, F., Bates, G.T. and Bartlein, P.J.
(1994) Lake-atmosphere feedbacks associated with
paleolakes Bonneville and Lahontan. Nature 263,
665–668.
Hostetler, S., Pisias, N. and Mix, A. (2006) Sensitivity of
Last Glacial Maximum climate to uncertainties in tropical and subtropical ocean temperatures. Quaternary
Science Reviews 25, 1168–1185.
Hubbell, S.P., He, F., Condit, R., Borda-de-Agua, L., et al.
(2008) How many tree species are there in the Amazon
and how many of them will go extinct? Proceedings of
the National Academy of Sciences 105, 11498–11504.
Hutton, J. (1795) Theory of the Earth; With Proofs and Illustrations. Creech, Edinburgh.
Imbrie, J. and Imbrie, K.P. (1979) Ice Ages. Solving the
Mystery. Macmillan, London.
Imbrie, J., Hays, J.D., Martinson, D.G., et al. (1984) The
orbital theory of Pleistocene climate: support from a
revised chronology of the marine δ18O record. In:
Berger A.L., Imbrie, J., Hays, J., et al. (eds) Milankovitch
and Climate Part 1. D. Reidel, Dordrecht, pp. 269–305.
IPCC (2001) Climate Change 2001: The Physical Science Basis.
Contribution of Working Group I to the Third Assessment
Report of the Intergovernmental Panel on Climate Change.
Cambridge University Press, Cambridge.
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.
Ivanochko, T.S., Ganeshram, R.S., Brummer, G.-J., et al.
(2005) Variations in tropical convection as an amplifier
of global climate change at the millennial scale. Earth
and Planetary Science Letters 235, 302–314.
Jackson, I.J. (1989) Climate, Water and Agriculture in the
Tropics. Longman, Harlow.
Jamieson, T.F. (1863) On the parallel roads of Glen Roy,
and their place in the history of the glacial period.
Introduction
Quarterly Journal of the Geological Society of London 19,
235–259.
Keppler, F., Hamilton, J.T., Braß M. and Rockmann, T.
(2006) Methane emissions from terrestrial plants under
aerobic conditions. Nature 439, 187–191.
Kerr, R.A. (2001) The tropics return to the climate system.
Science 292, 660–661.
Kohfeld, K.E. and Harrison, S.P. (2001) DIRTMAP: the
geological record of dust. Earth-Science Reviews 54,
81–114.
Köppen, W. (1936) Das geographische System de
Klimate. In: Handbuch der Klimatologie, Vol. 1. Borntager, Berlin.
Kottek, M., Grieser, J., Beck, C., et al. (2006) World Map
of the Köppen-Geiger climate classification updated.
Meteorologische Zeitschrift 15, 259–263.
Kroon, D., Steens, T. and Troelstra, S.R. (1991) Onset of
monsoonal related upwelling in the western Arabian
Sea as revealed by planktonic foraminifers. Proceedings
of the Ocean Drilling Program, Scientific results, vol. 117.
College Station Texas, ODP, pp. 257–263.
Kucera, M., Rosell-Melé, A., Schneider, R., et al. (2005)
Multiproxy approach for the reconstruction of glacial
ocean surface (MARGO). Quaternary Science Reviews 24,
814–819.
Kuhn, B.F. (1787) Versuch über den Mechanismus der
Gletscher. Magazin für die Naturkunde Helvetiens 1,
117–136.
Kutzbach, J.E. and Guetter, P.J. (1986) The influence of
changing orbital parameters and surface boundary conditions on climate simulations for the past 18,000 years.
Journal of Atmospheric Sciences 43, 1726–1759.
Larrasoaña, J.C., Roberts, A.P., Rohling, E.J., et al. (2003)
Three million years of monsoon variability over the
northern Sahara. Climate Dynamics 21, 689–698.
Lawrence, K.T., Liu, Z. and Herbert, T.D. (2006) Evolution
of the eastern tropical Pacific through Plio-Pleistocene
glaciation. Science 312, 79–83.
Lea, D.W., Pak, D.K., Peterson, J.C. and Hughen, K.A.
(2003) Synchroneity of tropical and high-latitude
Atlantic temperatures over the last glacial termination.
Science 301, 1361–1364.
Leng, M.J. (ed.) (2006) Isotopes in Palaeoenvironmental
Research. Developments in Paleoenvironmental Research,
Volume 10. Springer, Dordrecht.
Livingstone, D. (1857) Missionary Travels and Researches in
South Africa. Ward, Lock and Co., New York.
Loulergue, L., Schilt, A., Spahni, R., et al. (2008)
Orbital and millennial-scale features of atmospheric
methane over the past 800,000 years. Nature 354,
383–386.
29
Lourandos, H. (1997) Continent of Hunter-Gatherers: New
Perspectives in Australian Prehistory. Cambridge University Press, Cambridge.
Lowe, J.J. and Walker, M.J.C. (1997) Reconstructing Quaternary environments (2nd edition). Longman, Harlow.
Mace, G., Masundire, H., Baillie, J. coordinating lead
authors (2005) Biodiversity. In: Hassan, R., Scholes, R.
and Ash, N. (eds) Ecosystems and Human Well-being:
Current State and Trends, volume 1. Island Press, Washington, pp. 77–122.
Maher, B.A., Prospero, J.M., Mackie, D., et al. (2010)
Global connections between aeolian dust, climate
and ocean biogeochemistry at the present day and at
the last glacial maximum. Earth-Science Reviews 99,
61–97.
Manabe, S. and Hahn, D.G. 1977. Simulation of the tropical climate of an ice age. Journal of Geophysical Research
82, 3889–3911.
Martin, J.H., Gordon, R.M. and Fitzwater, S.E (1991)
The case for iron. Limnology and Oceanography 36,
1793–1802.
McGregor, G.R. and Nieuwolt, S. (1998) Tropical Climatology. John Wiley & Sons, Chichester.
McGuffie, K. and Henderson Sellers, A. (2001) Forty
years of numerical climate modelling. International
Journal of Climatology 21, 1067–1109.
Medina-Elizalde, M. and Lea, D.W. (2010) Late Pliocene
equatorial Pacific. Paleoceanography 25, PA2208, doi:
10.1029/2009PA001780.
Meigs, P. (1953) World distribution of arid and semi-arid
homoclimates. In: Arid Zone Hydrology. UNESCO Arid
Zone Research Series 1, 203–209.
Middleton, N.J., Thomas, D.S.G. and UNEP (eds) (1997)
World Atlas of Desertification (2nd edition). UNEP.
Milankovitch, M.M. (1941) Canon of Insolation and
the Ice Age Problem. Koniglich Serbische Akademie,
Belgrade.
Mix, A., Morey, A.E., Pisias, N.G. and Hostetler, S.W.
(1999) Foraminiferal faunal estimates of paleotemperature: circumventing the no-analog problem yields cool
ice-age tropics Paleoceanography 14, 350–359.
Molnar, P and Cane, M.A. (2002) El Niño’s tropical
climate and teleconnections as a blueprint for pre-Ice
Age climates. Paleoceanography 17, 1021, doi: 10.
1029/2001PA000663.
Mudelsee, M. and Raymo, M.E. (2005) Slow dynamics of
the Northern Hemisphere glaciations. Paleoceanography
20, PA4022, doi:10.1029/2005PA001153.
Munyikwa, K. (2005) Synchrony of Southern Hemisphere Late Pleistocene arid episodes: a review of luminescence chronologies from arid aeolian landscapes
30
Chapter 1
south of the Equator. Quaternary Science Reviews 24,
2555–2583.
Nash D.J. and Endfield, G.H. (2002) A nineteenth
century climate chronology for the Kalahari region of
central southern Africa derived from missionary correspondence. International Journal of Climatology 22,
821–841.
Nash, D.J. and Endfield, G.H. (2008) ‘Splendid rains have
fallen’: links between El Niño and rainfall variability in
the Kalahari, 1840–1900. Climatic Change 86, 257–290.
Nash, D.J. and Grab, S.W. (2010) ‘A sky of brass and
burning winds’: documentary evidence of rainfall
variability in Lesotho, 1824–1900. Climatic Change 101,
617–653.
Nederbragt, A.J. and Thurow, J. (2005) Amplitude of
ENSO cycles in the Santa Barbara Basin, off California,
during the past 15,000 years. Journal of Quaternary
Science 20, 447–456.
Newberry, J.S. (1861) Report upon the Colorado River of the
West, Part III. Geological report, Washington DC.
Nicholson, S.E. (2000) The nature of rainfall variability
over Africa on time scales of decades to millennia.
Global and Planetary Change 26, 137–158.
Nicholson, S.E. (2001) Climatic and environmental
change in Africa during the last two centuries. Climate
Research 17, 123–144.
Nott, J., Thomas, M.F. and Price, D.M. (2001) Alluvial
fans, landslides and Late Quaternary climatic change in
the wet tropics of northeast Queensland. Australian
Journal of Earth Sciences 48, 875–882.
O’Connell, J.F. and Allen, F.J. (2005) Dating the colonization of Sahul (Pleistocene Australia-New Guinea): a
review of recent research. Journal of Archaeological
Science 31, 835–853.
Oman, L., Robock, A., Stenchikov, G.L. and Thordarson,
T. (2006) High-latitude eruptions cast shadow over the
African monsoon and the flow of the Nile. Geophysical
Research Letters 33, L18711, doi:10.1029/2006GL
027665.
Ortlieb, L. (2000) The documentary historical records of
El Niño events in Peru: an update of the Quinn record
(sixteenth through nineteenth centuries). In: Diaz, H.
and Markgraf, V. (eds) El Niño and the Southern
Oscillation: Multiscale Variability and Global and Regional
Impacts. Cambridge University Press, Cambridge, pp.
207–295.
Paillard, D., Labeyrie, L. and Yiou, P. (1996) Macintosh
program performs time-series analysis. Eos Transactions
AGU 77, 379.
Parker, A.G., Eckersley, L., Smith, M.M., et al. (2004)
Holocene vegetation dynamics in the northeastern
Rub’ al-Khali desert, Arabian Peninsula: a phytolith,
pollen and carbon isotope study. Journal of Quaternary
Science 19, 665–676.
Patterson, L. J., Sturchio, N.C., Kennedy, B.M., et al.
(2005) Cosmogenic, radiogenic, and stable isotopic
constraints on groundwater residence time in the
Nubian Aquifer, Western Desert of Egypt, Geochemistry
Geophysics Geosystems 6, Q01005, doi:10.1029/
2004GC000779.
Pearson, S. and Dodson, J.R. (1993) Stick-nestrat middens
as sources of palaeoecological data in Australian deserts.
Quaternary Research 39, 347–354.
Peltier, W.R. and Fairbanks, R.G. (2008) Barbados Coral
Extended Last Glacial Maximum Sea Level Record.
IGBP PAGES/World Data Center for Paleoclimatology
Data Contribution Series # 2008-101. NOAA/NCDC
Paleoclimatology Program, Boulder CO, USA.
Pena, L.D., Cacho, I., Ferretti, P. and Hall, M.A. (2008) El
Nino-Southern Oscillation-like variability during glacial
terminations and interlatitudinal teleconnections. Paleoceanography 23, PA3101, doi:10.1029/2008PA001620.
Peterson, L.C. and Haug, G.H. (2006) Variability in the
mean latitude of the Atlantic Intertropical Convergence
Zone as recorded by riverine inputs of sediments to the
Cariaco Basin (Venezuela). Palaeogeography, Palaeoclimatology, Palaeoecology 234, 97–113.
Petit, J.R., Jouzel, J., Raynaud, D., et al. (1999) Climate
and atmospheric history of the past 420,000 years
from the Vostok Ice Core, Antarctica. Nature 399,
429–436.
Philander, S.G. and Fedorov, A.V. (2003) Role of tropics
in changing the response to Milankovitch forcing some
three million years ago. Paleoceanography 18, 1045,
doi:10.1029/2002PA000837.
Playfair, J. (1802) Illustrations of the Huttonian Theory of the
Earth. Creech, Edinburgh.
Prell, W.L. and Kutzbach, J.E. (1987) Monsoon variability
over the past 150,000 years. Journal of Geophysical
Research 92, D7, 8411–8425.
Prell, W.L. and Kutzbach, J.E. (1992) Sensitivity of the
Indian monsoon to forcing parameters and implications
for its evolution. Nature 360, 647–652.
Price, W.A. (1944) Greater American deserts. Texas
Academy of Science Proceedings and Transactions 27,
163–170.
Prinn, R.D., Cunnold, D., Rasmussen, R., et al. (1990)
Atmospheric emissions and trends of nitrous oxide
deduced from 10 years of Ale-gauge data. Journal of
Geophysical Research-Atmospheres 95, 18369–18385.
Quinn, W.H., Neal, V.T., and Antunez e Mayolo, S. (1987)
El Niño occurrences over the past four and half
Introduction
centuries. Journal of Geophysical Research 93, 14449–
14461.
Raymo, M.E. and Ruddiman, W.F. (1992) Tectonic forcing
of late Cenozoic climate. Nature 359, 117–122.
Raynaud, D., Chappellaz, J., Barnola, J.M., et al. (1988)
Climatic and CH4 cycle implications of glacial-interglacial
CH4 change in the Vostok ice core. Nature 333,
655–657.
Reading, A.J., Thompson, R.D. and Millington, A.C.
(1995) Humid Tropical Environments. Blackwell, Oxford.
Richards, P.W. (1952) The Tropical Rainforest. An Ecological
Study. Cambridge University Press, Cambridge.
Richey, J.E, Melack, J.M., Aufdenkampe, A.K., et al.
(2002) Outgassing from Amazonian rivers and wetlands as a large tropical source of atmospheric CO2.
Nature 416, 617–620.
Roberts, R.G., Jones, R. and Smith, M.A. (1990) Thermoluminescence dating of a 50,000 year old human
occupation site in northern Australia. Nature 345,
153–156.
Roberts, R.G., Jones, R. and Smith, M.A. (1993) Optical
dating at Deaf Adder Gorge, Northern Territory, indicates human occupation between 53,000 and 60,000
years ago. Australian Archaeology 37, 58–59.
Roberts, R.G., Jones, R., Spooner, N.A., et al. (1994) The
human colonisation of Australia: optical dates of 53,000
and 60,000 years bracket human arrival at Deaf Adder
Gorge, Northern Territory. Quaternary Science Reviews 13,
573–583.
Robock, A. (2000) Volcanic eruptions and climate. Reviews
of Geophysics 38, 191–219.
Robock, A. (2002) The climatic aftermath. Science 295,
1242–1244.
Rosenfeld, D., Kaufman, Y. and Koren, I. (2006) Switching cloud cover and dynamical regimes from open to
closed Benard cells in response to aerosols suppressing
precipitation. Atmospheric Chemistry and Physics 6,
2503–2511.
Ruddiman, W.F. (2003) The anthropogenic greenhouse
era began thousands of years ago. Climatic Change 61,
261–293.
Ruddimann, W.F. (2006) Orbital changes and climate.
Quaternary Science Reviews 25, 3092–3112.
Ruddiman, W.F. and Kutzbach, J.E. (1989) Forcing of
late Cenozoic northern hemisphere climate by plateau
uplift in southern Asia and the American west. Journal
of Geophysical Research 94, D15, 18409–18427.
Ruddiman, W.F. and Raymo, M.E. (1988) Northern hemisphere climate regimes during the past 3 Ma: possible
tectonic connections. Philosophical Transactions of the
Royal Society of London B318, 411–430.
31
Russell, I.C. (1885) Geological History of Lake Lahontan:
a Quaternary lake of northwestern Nevada. US Geological Survey Monograph 11. USGS.
Sarachik, E.S. and Cane, M.A. (2010) The El Niño-Southern
Oscillation Phenomenon. Cambridge University Press,
Cambridge.
Savage, J.M., Goldman, D.P., Janos, D.P., et al. (1982)
Ecological Aspects of Development in the Humid Tropics.
National Academy Press, Washington DC.
Scheffer, M., Carpenter, S., Foley, J.A., et al. (2001)
Catastrophic shifts in ecosystems. Nature 413, 591–
596.
Scheffer, M., Bascompte, J., Brock, W. A., et al. (2009)
Early-warning signals for critical transitions. Nature
461, 53–59.
Schilt, A., Baumgartner, M., Blunier, T., et al. (2010)
Glacial–interglacial and millennial-scale variations in
the atmospheric nitrous oxide concentration during the
last 800,000 years. Quaternary Science Reviews 29,
182–192.
Schmittner, A. and Galbraith, E.D. (2008) Glacial
greenhouse-gas fluctuations controlled by ocean circulation changes. Nature 456, 373–376.
Schwarz, E.H.L. (1923) The Kalahari or Thirstland Redemption. Oxford University Press, Cape Town.
Shackleton, N.J., Berger, A. and Peltier, W.R. (1990) An
alternative astronomical calilbration of the lower
Pleistocene timescale based on ODP site 677. Transactions of the Royal Society of Edinburgh, Earth Science 81,
251–261.
Shulmeister, J., Rodbell, D.T., Gagan, M.K. and Seltzer,
G.O. (2006) Inter-hemispheric linkages in climate
change: paleo-perspectives for future climate change.
Climates of the Past 2, 167–185.
Singarayer, J.S., Valdes, P.J., Friedlingstein, P., et al.
(2011) Late Holocene methane rise caused by orbitally
controlled increase in tropical sources. Nature 470, 82–
85. doi:10.1038/nature09739.
Singhvi, A.K. and Porat, N. (2008) Impact of luminescence dating on geomorphological and palaeoclimate
research in drylands. Boreas 536–558.
Smagorinsky, J. (1963) General circulation experiments
with the primitive equations. I. The basic experiment.
Monthly Weather Review 91, 99–164.
Sowers, T. (2009) Atmospheric methane isotope records
covering the Holocene period. Quaternary Science Reviews
29, 213–221.
Stager, J.C., Ryves, D., Cumming, B.F., et al. (2005) Solar
variability and the levels of Lake Victoria, East Africa,
during the last millennium. Journal of Paleolimnology 33,
243–251.
32
Chapter 1
Street, F.A. and Grove, A.T. (1976) Environmental and
climatic implications of Late Quaternary lake level fluctuations in Africa. Nature 261, 385–390.
Street-Perrott, F.A., Mitchell, J.F.B., Marchand, D.S. and
Brunner, J.S. (1979) Global maps of lake-level fluctuations since 30 000 yr. BP. Quaternary Research 12,
83–118.
Street-Perrott, F.A., Ficken, K.J., Huang, Y. and Eglinton,
G. (2004) Late Quaternary changes in carbon cycling
on Mt. Kenya, East Africa: an overview of the δ13C
record in lacustrine organic matter. Quaternary Science
Reviews 23, 861–879.
Telfer, M.W. and Thomas, D.S.G. (2006) Complex
Holocene lunette dune development, South Africa:
implications for palaeoclimate and models of pan
development in arid regions. Geology 34, 853–
856.
Therrell, M.D., Stahle, D.W., Ries, L.P. and Shugart, H.H.
(2006) Tree-ring reconstructed rainfall variability in
Zimbabwe. Climate Dynamics 26, 677–685.
Thomas, M.F. (2008) Understanding the impacts of Late
Quaternary climate change in tropical and sub-tropical
regions. Geomorphology 101, 146–158.
Thomas, M.F., Nott, J. and Price, D.M. (2001) Late Quaternary stream sedimentation in the humid tropics: a
review with new data from NE Queensland, Australia.
Geomorphology 39, 53–68.
Thomas, D.S.G., Knight, M. and Wiggs, G.F.S. (2005)
Remobilization of southern African desert dune systems
by twenty-first century global warming. Nature 435,
1218–1221.
Thomas, M.F., Nott, J., Murray, A.S. and Price, D.M.
(2007) Fluvial response to late Quaternary climate
change in NE Queensland, Australia. Palaeogeography,
Palaeoclimatology, Palaeoecology 251, 119–136.
Thornthwaite, C.W. (1948) An approach towards a
rational classification of climate. Geographical Review 38,
55–94.
Trauth, M.H., Deino, A.L., Bergner, A.G.N. and Strecker,
M.R. (2003) East African climate change and orbital
forcing during the last 175 kyr BP. Earth and Planetary
Science Letters 206, 297–313.
Tudhope, A.W., Chilcott, C.P., McCulloch, M.T., et al.
(2001) Variability in the El Niño-Southern Oscillation
through a glacial-interglacial cycle. Science 291,
1511–1517.
Van Loon, H., Meehl, G.A. and Arblaster, J.M. (2004) A
decadal solar effect in the tropics in July and August.
Journal of Atmospheric and Solar-Terrestrial Physics 66,
1767–1778.
Versteegh, G.J.M. (2005) Solar forcing of climate. 2:
evidence from the past. Space Science Reviews 120,
243–286.
Walker, G.T. (1924) Correlation in seasonal variations in
weather IV. A further study of world weather. Memoir
India Meteorological Department 24, 275–332.
Wang, B. and Ding, Q. (2008) Global monsoon: dominant
mode of annual variation in the tropics. Dynamics of
Atmospheres and Oceans 44, 165–183.
Wang, X., Auler, A.S., Edwards, R.L., et al. (2006)
Interhemispheric anti-phasing of rainfall during the
last glacial period. Quaternary Science Reviews 25,
3391–3403.
Wang, X., Auler, A.S., Edwards, R.L., et al. (2007) Millennial scale precipitation changes in southern Brazil
over the past 90,000 years. Geophysical Research Letters
34, L23701.
Wang, Y., Cheng, H., Edwards, R.L., et al. (2001) A
high resolution absolute-dated Late Pleistocene monsoon record from Hulu Cave, China. Science 294,
2345–2348.
Wang, Y., Cheng, H., Edwards, R.L., et al. (2008) Millennial- and orbital-scale changes in the East Asian
monsoon over the past 224,000 years. Nature 451,
1090–1093.
Williams, C., Flower, B.P., Hastings, D.W., et al. (2010)
Deglacial abrupt climate change in the Atlantic Warm
Pool: a Gulf of Mexico perspective. Paleoceanography 25,
PA4221, doi:10.1029/2010PA001928.
Williams, M.A., Ambrose, S.H., van der Kars, S., et al.
(2009) Environmental impact of the 73 ka Toba supereruption in South Asia. Palaeogeography, Palaeoclimatology, Palaeoecology 284, 295–314.
Wilson, E.O. (ed.) (1986) Biodiversity. National Academy
Press, Washington D.C.
Wolff, E. and Spahni, R. (2007) Methane and nitrous
oxide in the ice core record. Philosophical Transactions of
the Royal Society A 365, 1775–1792.
Yoon, S.C., Won, J.G., Omar, A.H., et al. (2005) Estimation of the radiative forcing by key aerosol types in
worldwide locations using a column model and
AERONET data. Atmospheric Environment 39, 6620–
6630.
Yue, X., Wang, H.J., Liao, H. and Jiang, D.B. (2011) Simulation of the direct radiative effect of mineral dust
aerosol on the climate at the Last Glacial Maximum.
Journal of Climate 24, 843–858.
Zachos, J., Pagani, M., Sloan, L., et al. (2001) Trends,
rhythms and aberrations in global climate 65Ma to
present. Science 292, 686–693.
Introduction
Zebiak, S.E. and Cane, M.A. (1987) A model El NiñoSouthern Oscillation. Monthly Weather Review 115,
2262–2278.
Zinke, J., Dullo, W.-Chr., Heiss, G.A. and Eisenhauer, A.
(2004) ENSO and Indian Ocean subtropical dipole variability is recorded in a coral record off southwest Mada-
33
gascar for the period 1659 to 1995. Earth and Planetary
Science Letters 228, 177–194.
Zinke, J., Pfeiffer, M., Timm, O., et al. (2005) Atmosphereocean dynamics in the Western Indian Ocean recorded
in corals. Philosophical Transactions of the Royal Society A
363, 121–142.