Historical climatology

Last updated
The 16th-century Skalholt Map of Norse America Skalholt-Karte.png
The 16th-century Skálholt Map of Norse America
One of Grimspound's hut circles Grimspound circle 1.jpg
One of Grimspound's hut circles

Historical climatology is the study of historical changes in climate and their effect on civilization from the emergence of homininis to the present day. This differs from paleoclimatology which encompasses climate change over the entire history of Earth. These historical impacts of climate change can improve human life and cause societies to flourish, or can be instrumental in civilization's societal collapse. The study seeks to define periods in human history where temperature or precipitation varied from what is observed in the present day.

Contents

The primary sources include written records such as sagas, chronicles, maps and local history literature as well as pictorial representations such as paintings, drawings and even rock art. The archaeological record is equally important in establishing evidence of settlement, water and land usage.

Techniques

In literate societies, historians may find written evidence of climatic variations over hundreds or thousands of years, such as phenological records of natural processes, for example viticultural records of grape harvest dates. In preliterate or non-literate societies, researchers must rely on other techniques to find evidence of historical climate differences.

Past population levels and habitable ranges of humans or plants and animals may be used to find evidence of past differences in climate for the region. Palynology, the study of pollens, can show not only the range of plants and to reconstruct possible ecology, but to estimate the amount of precipitation in a given time period, based on the abundance of pollen in that layer of sediment or ice. The distribution of diatoms in sediments can also be used to examine changes in salinity and climate over geologic eras. [1]

Role in human evolution

Changes in East African climate have been associated with the evolution of hominini. Researchers have proposed that the regional environment transitioned from humid jungle to more arid grasslands due to tectonic uplift [2] and changes in broader patterns of ocean and atmospheric circulation. [3] This environmental change is believed to have forced hominins to evolve for life in a savannah-type environment. Some data suggest that this environmental change caused the development of modern homimin features; however there exist other data that show that morphological changes in the earliest hominins occurred while the region was still forested. [4] Rapid tectonic uplift likely occurred in the early Pleistocene, [3] changing the local elevation and broadly reorganizing the regional patterns of atmospheric circulation. [5] [6] This can be correlated with the rapid hominin evolution of the Quaternary period. [2] Changes in climate at 2.8, 1.7, and 1.0 million years ago correlate well with observed transitions between recognized hominin species. [3] It is difficult to differentiate correlation from causality in these paleopanthropological and paleoclimatological reconstructions, so these results must be interpreted with caution and related to the appropriate time-scales and uncertainties. [7]

Ice ages

The eruption of the Toba supervolcano, 70,000 to 75,000 years ago reduced the average global temperature by 5 degrees Celsius for several years and may have triggered an ice age. It has been postulated that this created a bottleneck in human evolution. A much smaller but similar effect occurred after the eruption of Krakatoa in 1883, when global temperatures fell for about 5 years in a row.

Before the retreat of glaciers at the start of the Holocene (~9600 BC), ice sheets covered much of the northern latitudes and sea levels were much lower than they are today. The start of our present interglacial period appears to have helped spur the development of human civilization.

Role in human migration and agriculture

Climate change has been linked to human migration from as early as the end of the Pleistocene to the early twenty-first century. [8] [9] The effect of climate on available resources and living conditions such as food, water, and temperature drove the movement of populations and determined the ability for groups to begin a system of agriculture or continue a foraging lifestyle. [8]

Groups such as the inhabitants of northern Peru and central Chile, [10] the Saqqaq in Greenland, [11] nomadic Eurasian tribes in Historical China, [12] and the Natufian culture in the Levant all display migration reactions due to climatic change. [8]

Further descriptions of specific cases

In northern Peru and central Chile climate change is cited as the driving force in a series of migration patterns from about 15,000 B.C. to approximately 4,500 B.C. Between 11,800 B.C. and 10,500 B.C. evidence suggests seasonal migration from high to low elevation by the natives while conditions permitted a humid environment to persist in both areas. Around 9,000 B.C. the lakes that periodically served as a home to the natives dried up and were abandoned until 4,500 B.C. [10] This period of abandonment is a blank segment of the archeological record known in Spanish as the silencio arqueológico. During this break, there exists no evidence of activity by the natives in the lakes area. The correlation between climate and migratory patterns leads historians to believe the Central Chilean natives favored humid, low-elevation areas especially during periods of increased aridity. [10]

The different inhabitants of Greenland, specifically in the west, migrated primarily in response to temperature change. The Saqqaq people arrived in Greenland around 4,500 B.P. and experienced moderate temperature variation for the first 1,100 years of occupation; near 3,400 B.P. a cooling period began that pushed the Saqqaq toward the west. A similar temperature fluctuation occurred around 2,800 B.P. that led to the abandonment of the inhabited Saqqaq region; this temperature shift was a decrease in temperature of about 4 °C over 200 years. [11] Following the Saqqaq dominance, other groups such as the Dorset people inhabited west Greenland; the Dorset were sea-ice hunters that had tools adapted to the colder environment. The Dorset appeared to leave the region around 2,200 B.P. without clear connection to the changing environment. Following the Dorset occupation, the Norse began to appear around 1,100 B.P. in west Greenland during a significant warming period. [13] However, a sharp decrease in temperature beginning in 850 B.P. of about 4 °C in 80 years is thought to contribute to the demise of initial Norse occupation in western Greenland. [11]

In Historical China over the past 2,000 years, migration patterns have centered around precipitation change and temperature fluctuation. Pastoralists moved in order to feed the livestock that they cared for and to forage for themselves in more plentiful areas. [12] During dry periods or cooling periods the nomadic lifestyle became more prevalent because pastoralists were seeking more fertile ground. The precipitation was a more defining factor than temperature in terms of its effects on migration. The trend of the migrating Chinese showed that the northern pastoralists were more affected by the fluctuation in precipitation than the southern nomads. In a majority of cases, pastoralists migrated further southward during changes in precipitation. [12] These movements were not classified by one large event or a specific era of movement; rather, the relationship between climate and nomadic migration is relevant from "a long term perspective and on a large spatial scale." [12]

The Natufian population in the Levant was subject to two major climatic changes that influenced the development and separation of their culture. As a consequence of increased temperature, the expansion of the Mediterranean woodlands occurred approximately 13,000 years ago; with that expansion came a shift to sedentary foraging adopted by the surrounding population. [8] Thus, a migration toward the higher-elevation woodlands took place and remained constant for nearly 2,000 years. This era ended when the climate became more arid and the Mediterranean forest shrank 11,000 years ago. Upon this change, some of the Natufian populations nearest sustainable land transitioned into an agricultural way of life; sustainable land was primarily near water sources. Those groups that did not reside near a stable resource returned to the nomadic foraging that was prevalent prior to sedentary life. [8]

Historical and prehistoric societies

The rise and fall of societies have often been linked to environmental factors. [14]

Evidence of a warm climate in Europe, for example, comes from archaeological studies of settlement and farming in the Early Bronze Age at altitudes now beyond cultivation, such as Dartmoor, Exmoor, the Lake District and the Pennines in Great Britain. The climate appears to have deteriorated towards the Late Bronze Age however. Settlements and field boundaries have been found at high altitude in these areas, which are now wild and uninhabitable. Grimspound on Dartmoor is well preserved and shows the standing remains of an extensive settlement in a now inhospitable environment.

Some parts of the present Saharan desert may have been populated when the climate was cooler and wetter, judging by cave art and other signs of settlement in Prehistoric Central North Africa.

Societal growth and urbanization

Approximately one millennium after the 7 ka slowing of sea-level rise, many coastal urban centers rose to prominence around the world. [15] It has been hypothesized that this is correlated with the development of stable coastal environments and ecosystems and an increase in marine productivity (also related to an increase in temperatures), which would provide a food source for hierarchical urban societies. [15]

Societal collapse

The last written records of the Norse Greenlanders are from a 1408 marriage in Hvalsey Church -- today the best-preserved of the Norse ruins. Hvalsey Church.jpg
The last written records of the Norse Greenlanders are from a 1408 marriage in Hvalsey Church  — today the best-preserved of the Norse ruins.

Climate change has been associated with the historical collapse of civilizations, cities and dynasties. Notable examples of this include the Anasazi, [16] Classic Maya, [17] the Harappa, the Hittites, and Ancient Egypt. [18] Other, smaller communities such as the Viking settlement of Greenland [19] have also suffered collapse with climate change being a suggested contributory factor. [20]

There are two proposed methods of Classic Maya collapse: environmental and non-environmental. The environmental approach uses paleoclimatic evidence to show that movements in the Intertropical Convergence Zone likely caused severe, extended droughts during a few time periods at the end of the archaeological record for the classic Maya. [21] The non-environmental approach suggests that the collapse could be due to increasing class tensions associated with the building of monumental architecture and the corresponding decline of agriculture, [22] increased disease, [23] and increased internal warfare. [24]

The Harappa and Indus civilizations were affected by drought 4,500–3,500 years ago. A decline in rainfall in the Middle East and Northern India 3,800–2,500 is likely to have affected the Hittites and Ancient Egypt.

Medieval Warm Period

The Medieval Warm Period was a time of warm weather between about AD 800–1300, during the European Medieval period. Archaeological evidence supports studies of the Norse sagas which describe the settlement of Greenland in the 9th century AD of land now quite unsuitable for cultivation. For example, excavations at one settlement site have shown the presence of birch trees during the early Viking period. In the case of the Norse, the Medieval warm period was associated with the Norse age of exploration and Arctic colonization, and the later colder periods led to the decline of those colonies. [25] The same period records the discovery of an area called Vinland, probably in North America, which may also have been warmer than at present, judging by the alleged presence of grape vines.

Little Ice Age

Later examples include the Little Ice Age, well documented by paintings, documents (such as diaries) and events such as the River Thames frost fairs held on frozen lakes and rivers in the 17th and 18th centuries. The River Thames was made more narrow and flowed faster after old London Bridge was demolished in 1831, and the river was embanked in stages during the 19th century, both of which made the river less liable to freezing.

The Frozen Thames, 1677 The Frozen Thames 1677.jpg
The Frozen Thames, 1677

The Little Ice Age brought colder winters to parts of Europe and North America. In the mid-17th century, glaciers in the Swiss Alps advanced, gradually engulfing farms and crushing entire villages. The River Thames and the canals and rivers of the Netherlands often froze over during the winter, and people skated and even held frost fairs on the ice. The first Thames frost fair was in 1607; the last in 1814, although changes to the bridges and the addition of an embankment affected the river flow and depth, diminishing the possibility of freezes. The freeze of the Golden Horn and the southern section of the Bosphorus took place in 1622. In 1658, a Swedish army marched across the Great Belt to Denmark to invade Copenhagen. The Baltic Sea froze over, enabling sledge rides from Poland to Sweden, with seasonal inns built on the way. The winter of 1794/1795 was particularly harsh when the French invasion army under Pichegru could march on the frozen rivers of the Netherlands, while the Dutch fleet was fixed in the ice in Den Helder harbour. In the winter of 1780, New York Harbor froze, allowing people to walk from Manhattan to Staten Island. Sea ice surrounding Iceland extended for miles in every direction, closing that island's harbours to shipping.

The severe winters affected human life in ways large and small. The population of Iceland fell by half, but this was perhaps also due to fluorosis caused by the eruption of the volcano Laki in 1783. Iceland also suffered failures of cereal crops and people moved away from a grain-based diet. The Norse colonies in Greenland starved and vanished (by the 15th century) as crops failed and livestock could not be maintained through increasingly harsh winters, though Jared Diamond noted that they had exceeded the agricultural carrying capacity before then. In North America, American Indians formed leagues in response to food shortages. In Southern Europe, in Portugal, snow storms were much more frequent while today they are rare. There are reports of heavy snowfalls in the winters of 1665, 1744 and 1886.

In contrast to its uncertain beginning, there is a consensus that the Little Ice Age ended in the mid-19th century.

Evidence of anthropogenic climate change

Through deforestation and agriculture, some scientists have proposed a human component in some historical climatic changes. Human-started fires have been implicated in the transformation of much of Australia from grassland to desert. [26] If true, this would show that non-industrialized societies could have a role in influencing regional climate. Deforestation, desertification and the salinization of soils may have contributed to or caused other climatic changes throughout human history.

For a discussion of recent human involvement in climatic changes, see Attribution of recent climate change.

See also

Related Research Articles

The Holocene is the current geological epoch. It began approximately 9,700 years before the Common Era (BCE). It follows the Last Glacial Period, which concluded with the Holocene glacial retreat. The Holocene and the preceding Pleistocene together form the Quaternary period. The Holocene has been identified with the current warm period, known as MIS 1. It is considered by some to be an interglacial period within the Pleistocene Epoch, called the Flandrian interglacial.

<span class="mw-page-title-main">Pleistocene</span> First epoch of the Quaternary Period

The Pleistocene is the geological epoch that lasted from c. 2.58 million to 11,700 years ago, spanning the Earth's most recent period of repeated glaciations. Before a change was finally confirmed in 2009 by the International Union of Geological Sciences, the cutoff of the Pleistocene and the preceding Pliocene was regarded as being 1.806 million years Before Present (BP). Publications from earlier years may use either definition of the period. The end of the Pleistocene corresponds with the end of the last glacial period and also with the end of the Paleolithic age used in archaeology. The name is a combination of Ancient Greek πλεῖστος (pleîstos), meaning "most", and καινός, meaning "new".

<span class="mw-page-title-main">Little Ice Age</span> Climatic cooling after the Medieval Warm Period (16th–19th centuries)

The Little Ice Age (LIA) was a period of regional cooling, particularly pronounced in the North Atlantic region. It was not a true ice age of global extent. The term was introduced into scientific literature by François E. Matthes in 1939. The period has been conventionally defined as extending from the 16th to the 19th centuries, but some experts prefer an alternative timespan from about 1300 to about 1850.

<span class="mw-page-title-main">Climate variability and change</span> Change in the statistical distribution of climate elements for an extended period

Climate variability includes all the variations in the climate that last longer than individual weather events, whereas the term climate change only refers to those variations that persist for a longer period of time, typically decades or more. Climate change may refer to any time in Earth's history, but the term is now commonly used to describe contemporary climate change, often popularly referred to as global warming. Since the Industrial Revolution, the climate has increasingly been affected by human activities.

The Younger Dryas, which occurred circa 12,900 to 11,700 years BP, was a return to glacial conditions which temporarily reversed the gradual climatic warming after the Last Glacial Maximum, which lasted from circa 27,000 to 20,000 years BP. The Younger Dryas was the last stage of the Pleistocene epoch that spanned from 2,580,000 to 11,700 years BP and it preceded the current, warmer Holocene epoch. The Younger Dryas was the most severe and longest lasting of several interruptions to the warming of the Earth's climate, and it was preceded by the Late Glacial Interstadial, an interval of relative warmth that lasted from 14,670 to 12,900 BP.

<span class="mw-page-title-main">Toba catastrophe theory</span> Supereruption 74,000 years ago that may have caused a global volcanic winter

The Toba eruption was a supervolcano eruption that occurred around 74,000 years ago during the Late Pleistocene at the site of present-day Lake Toba in Sumatra, Indonesia. It is one of the largest known explosive eruptions in the Earth's history. The Toba catastrophe theory holds that this event caused a severe global volcanic winter of six to ten years and contributed to a 1,000-year-long cooling episode, leading to a genetic bottleneck in humans.

<span class="mw-page-title-main">Eemian</span> Interglacial period which began 130,000 years ago

The Eemian was the interglacial period which began about 130,000 years ago at the end of the Penultimate Glacial Period and ended about 115,000 years ago at the beginning of the Last Glacial Period. It corresponds to Marine Isotope Stage 5e. Although sometimes referred to as the "last interglacial", it was the second-to-latest interglacial period of the current Ice Age, the most recent being the Holocene which extends to the present day. The prevailing Eemian climate was, on average, around 1 to 2 degrees Celsius warmer than that of the Holocene. During the Eemian, the proportion of CO2 in the atmosphere was about 280 parts per million.

<span class="mw-page-title-main">Dansgaard–Oeschger event</span> Rapid climate fluctuation in the last glacial period

Dansgaard–Oeschger events, named after palaeoclimatologists Willi Dansgaard and Hans Oeschger, are rapid climate fluctuations that occurred 25 times during the last glacial period. Some scientists say that the events occur quasi-periodically with a recurrence time being a multiple of 1,470 years, but this is debated. The comparable climate cyclicity during the Holocene is referred to as Bond events.

<span class="mw-page-title-main">Global temperature record</span> Fluctuations of the Earths temperature over time

The global temperature record shows the fluctuations of the temperature of the atmosphere and the oceans through various spans of time. There are numerous estimates of temperatures since the end of the Pleistocene glaciation, particularly during the current Holocene epoch. Some temperature information is available through geologic evidence, going back millions of years. More recently, information from ice cores covers the period from 800,000 years before the present time until now. A study of the paleoclimate covers the time period from 12,000 years ago to the present. Tree rings and measurements from ice cores can give evidence about the global temperature from 1,000-2,000 years before the present until now. The most detailed information exists since 1850, when methodical thermometer-based records began. Modifications on the Stevenson-type screen were made for uniform instrument measurements around 1880.

The Holocene Climate Optimum (HCO) was a warm period in the first half of the Holocene epoch, that occurred in the interval roughly 9,500 to 5,500 years BP, with a thermal maximum around 8000 years BP. It has also been known by many other names, such as Altithermal, Climatic Optimum, Holocene Megathermal, Holocene Optimum, Holocene Thermal Maximum, Hypsithermal, and Mid-Holocene Warm Period.

<span class="mw-page-title-main">Abrupt climate change</span> Form of climate change

An abrupt climate change occurs when the climate system is forced to transition at a rate that is determined by the climate system energy-balance. The transition rate is more rapid than the rate of change of the external forcing, though it may include sudden forcing events such as meteorite impacts. Abrupt climate change therefore is a variation beyond the variability of a climate. Past events include the end of the Carboniferous Rainforest Collapse, Younger Dryas, Dansgaard–Oeschger events, Heinrich events and possibly also the Paleocene–Eocene Thermal Maximum. The term is also used within the context of climate change to describe sudden climate change that is detectable over the time-scale of a human lifetime, possibly as the result of feedback loops within the climate system or tipping points.

The Older Dryas was a stadial (cold) period between the Bølling and Allerød interstadials, about 14,000 years Before Present, towards the end of the Pleistocene. Its date range is not well defined, with estimates varying by 400 years, but its duration is agreed to have been around two centuries.

<span class="mw-page-title-main">Quaternary glaciation</span> Series of alternating glacial and interglacial periods

The Quaternary glaciation, also known as the Pleistocene glaciation, is an alternating series of glacial and interglacial periods during the Quaternary period that began 2.58 Ma and is ongoing. Although geologists describe this entire period up to the present as an "ice age", in popular culture this term usually refers to the most recent glacial period, or to the Pleistocene epoch in general. Since Earth still has polar ice sheets, geologists consider the Quaternary glaciation to be ongoing, though currently in an interglacial period.

<span class="mw-page-title-main">4.2-kiloyear event</span> Severe climatic event starting around 2200 BC

The 4.2-kiloyear BP aridification event, also known as the 4.2 ka event, was one of the most severe climatic events of the Holocene epoch. It defines the beginning of the current Meghalayan age in the Holocene epoch.

<span class="mw-page-title-main">Classic Maya collapse</span> Decline of classic Maya civilization

In archaeology, the classic Maya collapse is the decline of the Classic Maya civilization and the abandonment of Maya cities in the southern Maya lowlands of Mesoamerica between the 7th and 9th centuries. At Ceibal, the Preclassic Maya experienced a similar collapse in the 2nd century.

<span class="mw-page-title-main">8.2-kiloyear event</span> Rapid global cooling around 8,200 years ago

In climatology, the 8.2-kiloyear event was a sudden decrease in global temperatures that occurred approximately 8,200 years before the present, or c. 6,200 BC, and which lasted for the next two to four centuries. It defines the start of the Northgrippian age in the Holocene epoch. The cooling was significantly less pronounced than during the Younger Dryas cold period that preceded the beginning of the Holocene. During the event, atmospheric methane concentration decreased by 80 ppb, an emission reduction of 15%, by cooling and drying at a hemispheric scale.

<span class="mw-page-title-main">Bond event</span> North Atlantic ice rafting events

Bond events are North Atlantic ice rafting events that are tentatively linked to climate fluctuations in the Holocene. Eight such events have been identified. Bond events were previously believed to exhibit a roughly c. 1,500-year cycle, but the primary period of variability is now put at c. 1,000 years.

The Subatlantic is the current climatic age of the Holocene epoch. It started about 2,500 years BP and is still ongoing. Its average temperatures are slightly lower than during the preceding Subboreal and Atlantic. During its course, the temperature underwent several oscillations, which had a strong influence on fauna and flora and thus indirectly on the evolution of human civilizations. With intensifying industrialisation, human society started to stress the natural climatic cycles with increased greenhouse gas emissions.

<span class="mw-page-title-main">Roman Warm Period</span> Warm weather period, 250 BC to AD 400

The Roman Warm Period, or Roman Climatic Optimum, was a period of unusually-warm weather in Europe and the North Atlantic that ran from approximately 250 BC to AD 400. Theophrastus wrote that date trees could grow in Greece if they were planted but that they could not set fruit there. That is still the case today, which implies that South Aegean mean summer temperatures in the 4th and the 5th centuries BC were within a degree of modern ones. That and other literary fragments from the time confirm that the Greek climate was basically the same then as around 2000. Tree rings from the Italian Peninsula in the late 3rd century BC indicate a time of mild conditions there around the time of Hannibal's crossing of the Alps with imported elephants in 218 BC.

<span class="mw-page-title-main">Medieval Warm Period</span> Time of warm climate in the North Atlantic region lasting from c. 950 to c. 1250

The Medieval Warm Period (MWP), also known as the Medieval Climate Optimum or the Medieval Climatic Anomaly, was a time of warm climate in the North Atlantic region that lasted from c. 950 to c. 1250. Climate proxy records show peak warmth occurred at different times for different regions, which indicate that the MWP was not a globally uniform event. Some refer to the MWP as the Medieval Climatic Anomaly to emphasize that climatic effects other than temperature were also important.

References

  1. Fritz, S. C.; Juggins, S.; Battarbee, R. W.; Engstrom, D. R. (1991). "Reconstruction of past changes in salinity and climate using a diatom-based transfer function". Nature. 352 (6337): 706–708. Bibcode:1991Natur.352..706F. doi:10.1038/352706a0. ISSN   1476-4687. S2CID   4325091.
  2. 1 2 Gani, Nahid DS; Gani, M. Royhan; Abdelsalam, Mohamed G. (2007). "Blue Nile incision on the Ethiopian Plateau: Pulsed plateau growth, Pliocene uplift, and hominin evolution". GSA Today. 17 (9): 4. Bibcode:2007GSAT...17i...4G. doi: 10.1130/GSAT01709A.1 .
  3. 1 2 3 Demenocal, P. B. (1995). "Plio-Pleistocene African Climate" (PDF). Science. 270 (5233): 53–9. Bibcode:1995Sci...270...53D. doi:10.1126/science.270.5233.53. PMID   7569951. S2CID   617678. Archived from the original (PDF) on 2010-06-25.
  4. Winfried Henke, Ian Tattersall (eds.); in collaboration with Thorolf Hardt. (2007). Handbook of paleoanthropology. New York: Springer. ISBN   978-3-540-32474-4.{{cite book}}: CS1 maint: multiple names: authors list (link)
  5. Sepulchre, P; Ramstein, G; Fluteau, F; Schuster, M; Tiercelin, Jj; Brunet, M (Sep 2006). "Tectonic uplift and Eastern Africa aridification". Science. 313 (5792): 1419–23. Bibcode:2006Sci...313.1419S. doi:10.1126/science.1129158. ISSN   0036-8075. PMID   16960002. S2CID   4499083.
  6. Maslin, Ma; Christensen, B (Nov 2007). "Tectonics, orbital forcing, global climate change, and human evolution in Africa: introduction to the African paleoclimate special volume". Journal of Human Evolution. 53 (5): 443–64. doi:10.1016/j.jhevol.2007.06.005. ISSN   0047-2484. PMID   17915289.
  7. Behrensmeyer, Ak (Jan 2006). "Atmosphere. Climate change and human evolution". Science. 311 (5760): 476–8. doi:10.1126/science.1116051. ISSN   0036-8075. PMID   16439650. S2CID   128575322.
  8. 1 2 3 4 5 HENRY, DONALD O. (1989). From Foraging to Agriculture: The Levant at the End of the Ice Age . University of Pennsylvania Press. ISBN   9780812281378. JSTOR   j.ctv513bpq.
  9. Martiniello, Marco; Rath, Jan, eds. (2012). An Introduction to International Migration Studies: European Perspectives. Amsterdam University Press. doi:10.2307/j.ctt6wp6qz. ISBN   9789089644565. JSTOR   j.ctt6wp6qz.
  10. 1 2 3 Dillehay, Tom D. (2002). "Climate and Human Migrations". Science. 298 (5594): 764–765. doi:10.1126/science.1078163. JSTOR   3832641. PMID   12399573. S2CID   128971623.
  11. 1 2 3 D'Andrea, William J.; Huang, Yongsong; Fritz, Sherilyn C.; Anderson, N. John (2011). "Abrupt Holocene climate change as an important factor for human migration in West Greenland". Proceedings of the National Academy of Sciences of the United States of America. 108 (24): 9765–9769. Bibcode:2011PNAS..108.9765D. doi: 10.1073/pnas.1101708108 . JSTOR   25831309. PMC   3116382 . PMID   21628586.
  12. 1 2 3 4 Pei, Qing; Zhang, David D. (2014). "Long-term relationship between climate change and nomadic migration in historical China" (PDF). Ecology and Society. 19 (2). doi: 10.5751/ES-06528-190268 . JSTOR   26269570.
  13. Dugmore, Andrew J.; Keller, Christian; McGovern, Thomas H. (2007). "Norse Greenland Settlement: Reflections on Climate Change, Trade, and the Contrasting Fates of Human Settlements in the North Atlantic Islands". Arctic Anthropology. 44 (1): 12–36. doi:10.1353/arc.2011.0038. ISSN   1933-8139. PMID   21847839. S2CID   10030083.
  14. The Great Warming: Climate Change and the Rise and Fall of Civilizations. New York: Bloomsbury Press. 2008. ISBN   978-1-59691-392-9.
  15. 1 2 Day, John W.; Gunn, Joel D.; Folan, William J.; Yáñez-Arancibia, Alejandro; Horton, Benjamin P. (2007). "Emergence of Complex Societies After Sea Level Stabilized" (PDF). Eos, Transactions, American Geophysical Union . 88 (15): 169. Bibcode:2007EOSTr..88..169D. doi:10.1029/2007EO150001.
  16. Demenocal, P. B. (2001). "Cultural Responses to Climate Change During the Late Holocene" (PDF). Science . 292 (5517): 667–673. Bibcode:2001Sci...292..667D. doi:10.1126/science.1059827. PMID   11303088. S2CID   18642937.
  17. Hodell, David A.; Curtis, Jason H.; Brenner, Mark (1995). "Possible role of climate in the collapse of Classic Maya civilization". Nature. 375 (6530): 391. Bibcode:1995Natur.375..391H. doi:10.1038/375391a0. S2CID   4270939.
  18. Jonathan Cowie (2007). Climate change: biological and human aspects. Cambridge University Press. ISBN   9781107603561.
  19. transl. with introd. by Magnus Magnusson ... (1983). The Vinland sagas: the Norse discovery of America. Harmondsworth, Middlesex: Penguin Books. ISBN   978-0-14-044154-3.
  20. Diamond, Jared (2005). Collapse: How Societies Choose to Fail or Succeed. Viking Adult. ISBN   978-0-670-03337-9.
  21. Haug, Gh; Günther, D; Peterson, Lc; Sigman, Dm; Hughen, Ka; Aeschlimann, B (Mar 2003). "Climate and the collapse of Maya civilization". Science. 299 (5613): 1731–5. Bibcode:2003Sci...299.1731H. doi:10.1126/science.1080444. ISSN   0036-8075. PMID   12637744. S2CID   128596188.
  22. Hosler D, Sabloff JA, Runge D (1977). "Simulation model development: a case study of the Classic Maya collapse". In Hammond, Norman, Thompson, John L. (eds.). Social process in Maya prehistory: studies in honour of Sir Eric Thompson. Boston: Academic Press. ISBN   978-0-12-322050-9.
  23. Santley, Robert S.; Killion, Thomas W.; Lycett, Mark T. (Summer 1986). "On the Maya Collapse". Journal of Anthropological Research. 42 (2): 123–59. doi:10.1086/jar.42.2.3630485. S2CID   53000359.
  24. Foias, Antonia E.; Bishop, Ronald L. (1997). "Changing Ceramic Production and Exchange in the Petexbatun Region, Guatemala: Reconsidering the Classic Maya Collapse". Ancient Mesoamerica. 8 (2): 275. doi:10.1017/S0956536100001735. S2CID   162114230.
  25. Patterson, W.P.; Dietrich, K.A.; Holmden, C. (2007). Sea Ice and sagas: stable isotope evidence for two millennia of North Atlantic seasonality on the north Icelandic shelf. Arctic Natural Climate Change Workshop. Tromsø, Norway. CiteSeerX   10.1.1.132.4973 .
  26. Miller GH, Fogel ML, Magee JW, Gagan MK, Clarke SJ, Johnson BJ (July 2005). "Ecosystem Collapse in Pleistocene Australia and a Human Role in Megafaunal Extinction" (PDF). Science. 309 (5732): 287–290. Bibcode:2005Sci...309..287M. doi:10.1126/science.1111288. PMID   16002615. S2CID   22761857.

Further reading