You are on page 1of 9

Global temperature change during the Cretaceous, the Great

Ice Age, at present and in the future


Vitor Vieira Vasconcelos
PhD in Natural Sciences
Federal University of ABC
July 12th, 2016
Introduction
Current climate models shows that the accelerating trend of anthropogenic CO2 and
other greenhouse gases (GHG) emissions is a key factor for current global warming (IPCC, 2013,
p. 17). In order to understand the consequences of climate change, it becomes important to
model the climate of the previous geological eras, in order to understand Earths climatic
processes and better model future projections (HYDEN, COE & WILSON, 2007, p. 11 and 156).
The main objective of this essay is, by studying the Cretaceous and Great Ice Age environmental
conditions; analyse the likely future impacts of climate change.
The first section of this essay analyses the short, mid and long-term process of silicatecarbonate and organic carbon cycles in the Cretaceous, comparing with their respective
contemporary processes. The research about the greenhouse conditions faced during the
Cretaceous (from approximately 143 to 65 million years ago) may shed light on what could
happen if we maintain this trend of GHG emissions (SKELTON, 2006, p. 345).
Section 2 discusses the climate changes last 2.6 million years, denominated the Great
Ice Age. In this context, the section starts by exposing how the geological record can track past
temperature changes. In sequence, the rate and magnitude of temperature changes through
the Great Ice Age to modern times are discussed, and its effects on CO2 cycles, moonsons and
vegetation. Finally, the main characteristics of past, current and projected temperature changes
are analyzed, giving a clearer picture of the challenges and possibilities for climate change
modelling.

Section 1 Climate change in Cretaceous


CO2 Cycle in Cretaceous
The Cretaceous was market by a high ocean spreading rate, with constant and active
volcanic activity in mid ocean ridges, as well as flood basalts from continental rifting events, and
super-plumes that gave origin not only to volcanic islands chains but also to extensive basalt
plateaux under the oceans. The cumulative emission of CO2 from these volcanic activities
increased the greenhouse effect, increasing the global mean surface temperature (GMST).
Therefore, the high levels of CO2 in the atmosphere forced negative feedbacks in the carbon
biogeochemical cycle. One of these negative feedbacks was the increasing production of
1

carbonate platforms in low latitudes, which gave origin to the abundant Cretaceous limestone
in the geological record. The carbonate record of these negative feedbacks also includes the
chalk formed by the deposition of skeletons of planktonic organisms. Moreover, another
geological record of these environmental changes was the organic carbon burial as coal (in
continents, especially from the forests of the warm and wet high latitude areas) and kerogen (in
the oceans).
The silicate-carbonate cycle
In the short-term, the oceans absorb part of the increase in atmospheric CO2 (IPCC,
2013, p. 11). The phytoplankton uses this CO2 in their skeletons, and after death, their remains
precipitate and accumulate as chalk (SKELTON, 2006, p. 11). In the mid-term, as global
temperature increased in the Cretaceous, the ocean floor also became warmer (SKELTON, 2006,
p. 177), enabling chalk deposition in broader extensions without being re-dissolved during
precipitation.
The sea-level rise in the Cretaceous formed extensive shallow seas in low latitudes, while
higher evaporation released more CO2 in these saturated seas, allowing the development of
highly productive carbonate platform (SKELTON, 2006, p. 263). However, the chemical process
in carbonate platforms also releases part of the CO2 back into the atmosphere, as a mid-term
positive feedback in greenhouse conditions (SKELTON, 2006, p. 265).
In the long term, the carbonate from platforms and chalk acted as sinks to reduce
atmospheric CO2 (SKELTON, 2006, p. 270). However, any analogy with current age should be
cautions, because there are less shallow seas, and species of phytoplankton and carbonate
platforms are different from before.
The organic-carbon cycle
The Cretaceous world had warmer and wetter climate in high latitudes than today,
enabling better plant growth (SKELTON, 2006, p. 249). As a short-term effect, the thicker
vegetation decreased albedo, creating a positive feedback that increased temperatures even
more (SKELTON, 2006, p. 280). Plant growth takes CO2 from atmosphere through
photosynthesis, but the CO2 may come back through organic matter decomposition in the short
term. However, the period of relatively lower temperature in polar winter curtailed
decomposition (SKELTON, 2006, p. 161). In the mid-term, plant growth also accelerated
weathering and consequent sedimentation rates (SKELTON, 2006, p. 194), which was important
to cover the organic matter and store it as coal. In the long-term, coal storage in high latitudes
became a relevant carbon sink to gradually reduce atmospheric CO2 (SKELTON, 2006, p. 270).
As illustrated in the diagram of Figure 1, the increase in temperature and precipitation
due to the increase in atmospheric CO2 stimulates plant growth that, through photosynthesis,
sequester carbon from atmosphere and use it to store glucose and plant tissues. Depending on
the local environmental conditions, such as areas with high sediment deposition rates and
reduced organic decomposition (such as swamps and mires), a high share of this stored organic
carbon is buried and transformed into coal, removing the CO2 from the short-term carbon
2

atmospheric carbon cycle into the long term sink. Therefore, in the long-term, this carbon sink
of terrestrial organic carbon helped to partially mitigate greenhouse warming during the
Cretaceous, and, afterwards, especially towards the end of Cretaceous, to gradually return to
relatively lower CO2 levels.

Atmospher
ic CO2

Organic
Carbon
Burial

Climate
(temperatu
re and
precipitatio

Plant
photosynthe
sis and

Figure 1 Diagram illustrating the mean Earth systems feedbacks related to


atmospheric CO2 and organic carbon burial. Green arrows are positive feedbacks and
white arrows are negative feedbacks.
The high latitude organic deposition environments in the Cretaceous were formed under
warmer and year-round wet climatic conditions. These environments originated coal layers that
typically form in mires, that is, wet terrestrial environments where organic matter is deposited
in higher rates than siliciclastic sediments. These mires could be swamps (low-lying depressions
filled with water and hygrophyte plants), or in raised bogs, where the saturated soils was
maintained only by the local rainfall. In order to cope with the artic light regime, most of the
plants acquired deciduous behaviour, losing their leaves during the winter period of continuous
darkness, in order to avoid respiratory drain of the plant resources during that time. An even
more extreme adaptation would be the winter dieback of vegetation tissues, such as twigs or
even stalks, in order to regrow in the following spring. The lower temperature of the winter
helped to limit the decay of the fallen vegetation tissues, while the active sedimentation rates
of deltaic environments helped covering and compacting the organic layers.
Current greenhouse effect is increasing temperature in polar regions in a higher rate
than global mean temperature (IPCC, 2013, p. 20). This process is leading to melting of glaciers
and permafrost (IPCC, 2013, p. 24). These environmental changes will allow higher vegetation
productivity, possibly implying similar processes for the organic carbon cycle such as seen during
the Cretaceous.

The release of anthropogenic CO2 in the atmosphere, with its subsequent increase in
global warming, would cause significant impacts on polar regions. The first change, caused by
temperature warming, would be melting the glaciers and permafrost soils. These changes,
together with a higher concentration of atmospheric CO2, would allow higher rates of vegetation
growth in these areas. This vegetation would have a low albedo, absorbing more solar heat and
increasing even more the local temperature. Many plants and animals adapted to current polar
environment conditions would face risk of extinction with the new environmental changes.

Section 2 Climate change in the Great Ice Age


Temperature changes from geological records
Drilled ice cores from polar ice sheet are important proxies for temperature changes
during the Great Ice Age. As snow seasonally accumulated and was compressed into ice, it lets
a record of how the climate was during the deposition time of each layer. Using isotope analysis
of 18O and D, it is possible to model the local temperature changes at these drilling sites
(HYDEN, COE & WILSON, 2007, p. 115-116, Fig. 5.12), because cooler climate favors a
proportional deposition of lighter isotopes. In addition, the content CO2 and methane (CH4)
through these cores reflect the past global atmospheric composition (HYDEN, COE & WILSON,
2007, p. 121, fig. 5.16) of the two most important greenhouse gases (IPCC, 2013, 14), enabling
to infer the past greenhouse effect.
In addition, selected drilling sites at deep ocean floor provide additional 18O proxies
from foraminiferid fossils (HYDEN, COE & WILSON, 2007,, p. 84-89). The 18O of benthic
foraminiferids is a proxy of the extension of ice sheets, while the 18O of planktonic
foraminiferids also receives the influence of past sea surface temperature. Analysing both
records (benthic and planktonic), it is possible to separate the two signals of surface
temperature and ice sheet growth (HYDEN, COE & WILSON, 2007, p. 86-87). Dust records in
deep ocean and ice sheet cores are also a good proxy for temperature change, because the
winds are stronger and dustier during glacial periods (HYDEN, COE & WILSON, 2007, p. 119).
The study of pollen and beetle fossils in diverse drilling sites around the world are also
a good sedimentary proxy of temperature change. The temperature change can be inferred from
the variation in the record of species adapted to distinct climates (HYDEN, COE & WILSON, 2007,
p. 184-191). Using fossils collected in different sites, it is possible to model their migration
patterns as temperature changed (HYDEN, COE & WILSON, 2007, p. 187), i.e., developing a
spatial-temporal inference of past temperatures.
Rates and magnitudes of temperature change during the last 2.6 million years of Earth History
The Great Ice Age has faced many oscillations in temperature. A usual pattern from
sedimentary proxies is that the end of a glacial period would be followed by a fast temperature
elevation into a warmer interglacial period. In sequence, after some thousand years, the
interglacial would start a period of gradual cooling, afterwards followed by a fast cooling that
4

would lead to the glacial maximum (HYDEN, COE & WILSON, 2007, p. 92). However, this broad
pattern also comprises many smaller-scale oscillations, such as 6-20 thousand years Bond Cycles
(HYDEN, COE & WILSON, 2007, p. 137), which depict gradual cooling periods. Bond Cycles are
still composed of even smaller scale oscillations called Dansgaard-Oeschger (D-O) events
(HYDEN, COE & WILSON, 2007, p. 136). The D-O oscillations define stadials (relatively cooler
periods) and interstadials (warmer periods).
The temperature changes during the Great Ice Age are in broad agreement with the
Milankovich cycles of 100, 41, 23 and 19 thousand years. Through these cycles, the strength of
summer insolation on the Northern Hemisphere is affected by changes in Earths orbital patterns
of eccentricity, obliquity and precession (HYDEN, COE & WILSON, 2007, p. 77 and 98). Cooler
summers would not be able to melt the winter snowfall, thereby increasing the thickness and
extension of ice sheets. Intriguingly, around 1 million years ago, the most significant pattern of
glaciation cycles changed from 41 to 100 thousand years (HYDEN, COE & WILSON, 2007, p. 98).
Models based on D changes in Antarctic ice cores estimate that Antarctic temperature
were 4 C lower than today in the last glacial maximum (18 thousand years ago), and 2 oC higher
during the last interglacial, 125 thousand years ago (HYDEN, COE & WILSON, 2007, p. 117). In
Northern Hemisphere, beetle records show a variation of almost 20 oC in Britain, from the last
glaciation maximum to early Holocene (8 thousand years ago) (HYDEN, COE & WILSON, 2007, p.
191).
o

Comparison of temperature change: past, present and future


Since the end of the Cretaceous period, 65.5 million years ago, Earth has been in a
process of gradual cooling (HYDEN, COE & WILSON, 2007, p. 93 and 156), also reflected by a
decreasing trend in atmospheric CO2 levels. As commented in the previous section, this gradual
cooling trend included many oscillations, paced by Milankovich Cycles. In accordance to these
cycles, and comparing the time extension of previous interglacials, a new period of glaciation
would start in the near future (HYDEN, COE & WILSON, 2007, p. 175).
However, anthropogenic emissions of GHG have been changing the natural trends of
temperature change. For example, atmospheric CO2 and CH4 concentrations are higher than in
any recorded moment of the last 800 thousand years (IPCC, 2013, p. 11). The correlation of CO2
and temperature through the Great Ice Age supports climate models assumptions that these
anthropogenic effects would lead to global warming (IPCC, 2013, p. 26).
It is also worth mentioning that the temperature stability during the Holocene, in the
last 10 thousand years, is an exception, rather than a rule, when compared to the frequent
oscillations during the Great Ice Age (HYDEN, COE & WILSON, 2007,, p. 143). Even the last
interglacial period (Eemian) had relevant cooling episodes, detected from evidences in deep
ocean foraminiferid fossils and pollen evidence in continental lands (HYDEN, COE & WILSON,
2007, p. 144).

Changes in CO2 cycle


Comparing CO2 and temperature proxies in Antarctic and Greenland ice cores,
atmospheric CO2 increased during interstadial and interglacial periods, whereas it decreased
during stadial/glacial periods. For example, atmospheric CO2 levels were around 300 ppm in the
last interglacial, but were just 200 ppm in the last glacial maximum. The ocean is the largest
reservoir that can store these changes in atmospheric CO2 at Milankovich-scales of climate
change.
During glacial periods, the lower temperature in the oceans increased its CO2 solubility,
enabling its absorption from the atmosphere. In addition, the stronger glacial winds increased
the vigour of oceanic upwelling systems, bringing more nutrient to surface and then enhancing
planktonic carbon pump through the precipitation of their shells (Figure 2).

Figure 2 Ocean circulation affecting CO2 cycle during the Great Ice Age
The chemical reaction of carbonate weathering and deposition is neutral in the total
sum, but these two processes where not in balance during the transition between glacials and
interglacials. For example, eustatic sea level fall during global cooling would expose carbonate
platforms on the continental shelf, and their weathering would bring CO2 from the atmosphere
to the oceans. During the transition from glacial/stadial to interglacial/interstadial, the relative
sea level rise stimulated chemical carbon deposition by coral reefs, but in their respective
chemical reaction, they pump part of the CO2 into the atmosphere again, reinforcing the
greenhouse trend in the short term.
In the long term, during the Great Ice Age, the weathering of silicate minerals, especially
from the uplifted ridges of Himalaya and Western USA, brought to oceans a great amount of
atmospheric CO2.
Changes on Moonsons
These climatic oscillations during the Great Ice Age caused changes in Earths
temperature, insolation and other climatic attributes. Particularly, summer monsoon strength

decreased during glacial periods, and increased during hotter periods (interglacials or
interstadials).
With lower temperature during glaciation, the continental air is less able to rise and to
produce a low-pressure zone that would attract the oceanic wet winds. In addition, with lower
temperatures in the oceans, there is less evaporation from the sea and thus less rainfall when
the summer ocean monsoon reach the continents. During the last glaciation maximum, eighteen
thousand years ago, there is sedimentary evidence that the Sahara region and Arabian Peninsula
had much broader sand dune deserts, because the monsoon rainfall was not strong enough to
reach these areas and support vegetation growth.
On the other hand, during interglacial periods, the opposite trend happens. The warm
continents heat the air, which rises, creating higher contrast in relation to the air pressure over
the oceans, thus attracting the humid winds to the continents. In the Holocene climatic
optimum, between nine and six thousand years ago, when the temperature used to be 1C
higher than now, the monsoon was so stronger that, in Africa, the Sahara was able to support
many permanent lakes and vegetation covered many sand dune areas.
These variations of the monsoon can be clearly observed using the proxy of the grain
size in loess-soil deposits from Tibetan Plateau and the methane concentration in glaciers ice
cores. The graphs of these proxies show an oscillatory pattern of monsoonal strength, in fairly
good correlation with the temperature changes recorded by 18O and D proxies in these ice
cores.
The couplets of loess and soil deposits in the Tibetan plateau give a proxy sedimentary
evidence for monsoon changes during the last 2.5 million years. The soil layers with low
carbonate content developed when the summer monsoon was strong and brought enough
rainfall to support vegetation growth. On the other hand, during glacial periods, cold and strong
winds from continent to sea deposited loess with higher carbonate content, which came from
the dryer deserts. The grain size in these sequences can be used as a quantitative variable,
assuming that increases in grain size indicate a relative increase in aridity.
Another important proxy would be the concentration of methane (CH4) in ice cores from
Antarctica (last 400 thousand years) and Greenland (last 110 thousand years, but with more
detailed time scale). As the methane was trapped in the ice from atmosphere, it gives a good
picture of global changes, because the mixing time in atmosphere is very rapid (less than 10
years) in geological time scale. The atmospheric methane signal responds strongly to the
productivity of the wetlands in low latitudes, because these wetlands thrive with stronger
monsoonal rainfall. However, the signal of atmospheric methane concentration can be
influenced by high latitude wetlands during interglacial periods, and can also be rapidly released
during glacial terminations from the methane clathrates on seabed.
Effects on Vegetation
In low latitudes, the main change in the transition from interglacials to glacials would
probably be the decrease in rainfall. This decrease would restrain the areas that support tropical
7

rainforest, which needs precipitation of at least 1,500mm/year. The reduction in the


concentration of atmospheric CO2 would increase plant evapotranspiration, with and effect
equivalent to reducing rainfall. The tropical rainforest would be fragmented into refuges,
surrounded by deciduous forests in South America and Savannah in Africa. In Africa, the
decrease in monsoonal rains would cause the Sahara desert to expand during glacials, pushing
the layers of savannah and forest farther south. During the transition from glacials to
interglacials, conversely, the tropical rainforests would expand again, and the Sahara desert
would shrink, becoming spotted with permanent lakes that supported vegetated ecosystems.
In high latitudes and high altitudes, temperature change was the main factor controlling
terrestrial vegetation patterns during these climatic transitions. In the transition to interglacials,
as temperature increased and the ice sheets and permafrost soils melted in northern
hemisphere, the vegetation migrated North again: first the tundra, then pine trees and, when
feasible, the temperate forest. The herbaceous species were able to migrate faster on the
exposed bareland of the previous icesheets, while forest migrated slower, taking centuries or
thousands of years until they reached the ecosystem climax. On the other hand, when the glacial
periods returned, the artic vegetation had to migrate to southern refuges again. In high
mountains, the stratified layers of glaciers, paramos, conifers and forest (from the upper layer
to the bottom layer, respectively) moved up during interglacials, and descended during glacials.
The top of the higher mountains, during interglacials, become a refuge of vegetation adapted to
colder climates, and this vegetation would spread again to lower elevations during glacial
periods.

Conclusion
A future world with increasing emissions of GHG adds a new factor that was not present
during most of the Great Ice Age. The climate models agree about the global warming trend in
the short-term, but there are still great uncertainties about longer-term climate projections,
because the future scenarios are different from our past records. Following this emissions trend,
human society is risking ruining the climate equilibrium that helped humanity to develop during
the current interglacial. Even if our society stops the GHG emissions, the feedback mechanisms
of the biochemical cycles would take centuries, thousands of even millions of years (i.e.,
geological timescale), to return to a new equilibrium (HYDEN, COE & WILSON, 2007, p. 209; IPCC,
2013, p. 28).
Geological research about past climates reveals many impacts that current society may
face if maintains the trend of GHG emissions, approaching greenhouse conditions closer to the
Cretaceous (SKELTON, 2006, p. 186). These geological evidences are especially important to
corroborate projections from climate models (IPCC, 2013, p. 26). Among many foreseen impacts,
there are higher temperature (IPCC, 2013, p. 20), sea-level rise (SKELTON, 2006, p. 69; IPCC,
2013, p. 25), increase in storm intensity (SKELTON, 2006, p. 179), decrease in polar albedo
(SKELTON, 2006, p. 280), loss of polar ice, as well as the other changes in organic and silicatecarbonate carbon cycles discussed along this essay. All these environmental changes may cause
8

an accelerated rate of extinction of species that fail to adapt to these new conditions. Moreover,
human adaption to these changes would cause high economic costs. A higher awareness of
these possible impacts may help our society to change these drivers of climate change.
References:
HYDEN, F.M.; COE, A.L.; WILSON, C.L. (2007) The Great Ice Age. Milton Keynes, The Open
University.
IPCC, 2013 (2013) Summary for Policymakers in Climate Change 2013: The Physical Science
Basis. Contribution of Working Group I to the Fifth Assessment Report of the Intergovernmental
Panel
on
Climate
Change
[Online].
Available
at
http://www.climatechange2013.org/images/report/WG1AR5_SPM_FINAL.pdf (Accessed 3
March 2016).
SKELTON, P. (ed.) (2006) The Cretaceous World. Cambridge, Cambridge University Press / Milton
Keynes, The Open University.

You might also like