Skip to main content

The Climate Downturn of 536–50

  • Chapter
  • First Online:
The Palgrave Handbook of Climate History

Abstract

This chapter surveys the evolution of research on the 536–50 CE climatic downturn and its human impacts. It presents the written evidence for atmospheric anomalies over the Mediterranean alongside the ever-growing wealth of relevant ice core and tree ring scholarship, and it highlights changes in reconstruction and interpretation as scholars reworked old evidence and injected new data. Judgments about the downturn’s historical significance in multiple world regions are discussed but not assessed in depth. In line with current evidence, the chapter concludes that the anomaly was a discontinuous complex of phenomena whose effects were extreme but varied across space and time. A cluster of very large volcanic eruptions triggered exceptional cooling and possibly drought across several parts of the globe. This was not a “536 event” or a “mystery cloud” of 12 or 18 months’ duration. It was a decade and a half of marked cold, with troughs in summer temperatures around 536, 540–1, and 545–6.

This is a preview of subscription content, log in via an institution to check access.

Access this chapter

Institutional subscriptions

Notes

  1. 1.

    A handful of antiquarians and Byzantinists drew attention to accounts of a c. 536 Mediterranean mystery clouding before the 1980s (Stathakopoulos, 2003, 247–49), but none envisioned this atmospheric phenomenon was part of a European, Eurasian, hemispheric, or global climatic event before NASA scientists Stothers and Rampino: Stothers and Rampino, 1983a, 412, 1983b, 6357, 6362–63, 6367, 6369; Stothers, 1984; Rampino, 1988, 87–88. Early Byzantinist scholarship notably includes Koder, 1996, and Farquharson, 1996, 266–68, 76–77.

  2. 2.

    Masson-Delmotte et al., 2014; Steig et al., 2013, 373; PAGES 2K Consortium, 2013, Tab. 1, Fig. 2; Jones et al., 2009, 6, 7. Although there remain many large gaps in our knowledge, limited evidence indicates temperature was not unusual in the mid-sixth century near the South Pole. Recent simulations of the climate forcing of downturn volcanism also suggest that the Southern Hemisphere was relatively unaffected: Toohey et al., 2016, 406. It is notable that Tambora too appears not to have much disturbed extratropical climates south of the equator: Raible et al., 2016, 569, 572, 582. The climate forcing of that 1815 eruption was slightly less than that of the c. 540 event: Sigl et al., 2015, 547–48, Extended Data Tab. 4. Yet, as Raible et al., 2016, remark (576), a dearth of climate records in the Southern Hemisphere may account for Tambora’s poor showing in the south. Of course, there are even fewer records for the sixth century.

  3. 3.

    Sigl et al., 2015, 547–48, Figs. 2 and 3, Extended Data Tabs. 4 and 5.

  4. 4.

    Büntgen et al., 2016. This LALIA falls within a longer period of less extreme cooling (known by many names, including Vandal Minimum, Late Roman Cold Period, Migration Period Pessimum, and Early Medieval Cold Anomaly) that commenced, depending on the proxy employed, in the fourth or fifth century and petered out in the seventh or eighth century. For example, Büntgen et al., 2011, 581; McCormick et al., 2012, 191–99.

  5. 5.

    Luterbacher et al., 2016, Fig. 1.

  6. 6.

    Toohey et al., 2016, 401, 405, 406, 410, Fig. 2.

  7. 7.

    Some historians have over-generalized the fragility of paleoclimate dating: try Moorhead, 2001, 143. The dendroclimatological data has proven robust. The ice core data is trickier. Yet the former cannot be problematized on account of the challenges the latter can present.

  8. 8.

    Bondesson and Bondesson, 2014, 63, for instance, claimed the cause of the downturn, which they consider both vast and severe, “remains unclear,” and they seem to suggest the event was restricted to the mid-530s, even though its volcanic origin was reconfirmed in 2008 (and only since reinforced) and its decadal duration was made evident no later than 1994.

  9. 9.

    The notable exception is Arjava, 2005, 73–94. Many in the humanities continue to read the paleoscience through Arjava’s paper, though much has changed since 2005. See Power, 2012, 190; Lee, 2013, 290.

  10. 10.

    Keys, 2000; Wickham, 2005, 549.

  11. 11.

    McCormick writes of a “tremendous volcanic winter” in 536 with widespread atmospheric effects that “must have had serious economic and human consequences” but which only “weakened and did not destroy” the Roman Empire revived under Justinian: McCormick, 2013, 72, 88.

  12. 12.

    Cassiodorus’ first appearance: Rampino, 1988, 87.

  13. 13.

    Cameron, 1985, 14.

  14. 14.

    Procopius, 1916, IV.14, 328–29.

  15. 15.

    Cassiodorus, Variae 12.25, 518–20.

  16. 16.

    Lydian, 1897, 25. On the misdating: Arjava, 2005, 80.

  17. 17.

    Pseudo-Dionysius of Tel-Mahre, Chronicle, 65.

  18. 18.

    Michael the Syrian, 1901, 220–21.

  19. 19.

    Pseudo-Zachariah Rhetor, Chronicle 9.19, 370.

  20. 20.

    Pseudo-Zachariah Rhetor, Chronicle, 370 n. 305.

  21. 21.

    Arjava, 2005, 79.

  22. 22.

    Pseudo-Zachariah Rhetor, Chronicle 10.1, 399.

  23. 23.

    Marcellinus Comes, Chronicle, 39.

  24. 24.

    Arjava, 2005, 80–83; Stothers and Rampino, 1983b, 6362.

  25. 25.

    Notably: Stothers, 1984, 344–45; Rampino, 1988, 87: “the densest and most persistent dry fog in recorded history was observed during AD 536–537.”

  26. 26.

    Lydian, 1897, 25; Arjava, 2005, 80.

  27. 27.

    By which John may have meant Ethiopia or southern Arabia. Sixth-century Byzantines sometimes confused the two: Sarris, 2002, 171; Schneider, 2015, 184–202.

  28. 28.

    It was suggested John borrowed his explanation from Campestris who lived centuries earlier. Arjava thinks this dubious: Arjava, 2005, 81.

  29. 29.

    Keys, 2000, 253; Abbott et al., 2014b, 413.

  30. 30.

    Weisburd, 1985, 91–94; Houston, 2000, 71, 73, 77. Whether there is textual evidence for exceptional cold and drought in West Asia and Europe in the 540s remains to be determined. Previous searches have centered on 536.

  31. 31.

    Aston, 1896, 34–35.

  32. 32.

    Shultz, 2012, 122–24. There appears to be nothing potentially related to the downturn in Paekche Annals of the Samguk Sagi.

  33. 33.

    Koguryo Annals of the Samguk Sagi, 168–69. There appears to be nothing potentially related to the downturn in the Paekche Annals of the Samguk Sagi.

  34. 34.

    LaMarche and Hirschboeck, 1984, 121–26 (cf. Parker, 1985); Briffa, 2000, 87–105; Gao et al., 2008; Cole-Dai, 2010, 824–39.

  35. 35.

    Churakova et al., 2014, 145–49.

  36. 36.

    Esper et al., 2013, 2, 2015.

  37. 37.

    On these issues: Esper et al., 2015, 62–70; García-Suárez et al., 2009, 183–98.

  38. 38.

    535 registered as the second coldest June–January in an early TRW study of a Sierra Nevadan pine chronology spanning 1–1980 ce (536 placed first), TRW and cell wall thickness analysis also drew attention to a 532 cold plunge in the aforementioned Altai series, and TRW analysis of the associated Sakha series revealed a pre-downturn 533 low. These Russian lows may be connected to local volcanism and suggest that the downturn had an early start in Siberia.

  39. 39.

    Cook et al., 2015.

  40. 40.

    Büntgen et al., 2011.

  41. 41.

    For instance: Esper et al., 2013, 736, Fig. 3.

  42. 42.

    Cook et al., 2006, 689–99; Larsen et al., 2008.

  43. 43.

    Pearson et al., 2012.

  44. 44.

    Major low-latitude eruptions, like the c. 540 event, are known to reduce global mean precipitation: Iles et al., 2013. Fischer et al., 2007, finds drier conditions in Central and Eastern Europe after more recent large (tropical) eruptions. Also Luterbacher and Pfister, 2015.

  45. 45.

    Pearson et al., 2012, 3405, 3411–12. Vesuvius’ 472 eruption also does not register in this series. Narrow rings are apparent, however, at the 475 mark (see below), perhaps indicating a post-472 eruption dry spell.

  46. 46.

    Esper et al., 2013, 736, Fig. 3.

  47. 47.

    See Fig. 2.6 and references there cited in Luterbacher et al., 2012, 103.

  48. 48.

    Tan et al., 2003, 1617; Zhang et al., 2008, 940, 941 (Fig. 1).

  49. 49.

    Holzhauser et al., 2005; Thompson et al., 1985, 973, 1994, 85, 87, 92. The second study indicates dryness recommencing c. 570, following a decade-long hiatus, and continuing until 610. Chu et al., 2011, 789–90; Van Bellen et al., 2015, 1, 9. The Patagonian dry period, which seems to predate but span the downturn, is visible as well in another southern South American proxy too: Moreno et al., 2014.

  50. 50.

    Kobashi et al., 2011.

  51. 51.

    See note 2 above.

  52. 52.

    Sixth-century sections of long high-resolution Central American proxies are wanting. The region is held to suffer heightened aridity following large eruptions—see Gill and Keating, 2002, 125–33.

  53. 53.

    Hodell et al., 1995, 393 (Fig. 3); Curtis et al., 1996, 41, 44–46; Hodell et al., 2001, 1368 (Fig. 2), 2005, 1421, 1424 (Figs. 10, 15). These studies focus on the more severe and prolonged droughts corresponding to the classical “collapse,” not the hiatus, though the latter is visible in them. The very existence of severe and prolonged classical-era droughts, however, has been questioned. The Chichancanab data has been reassessed and it has been argued that the arid cycles identified in the aforementioned 2001 and 2005 papers are “methodological artifacts”: Carleton et al., 2014, 151–61. Dry conditions evident in the Chichancanab data, though, appear in other independent proxies from the region: Wahl et al., 2014, 23.

  54. 54.

    Medina-Elizalde et al., 2010, 260 (Fig. 7).

  55. 55.

    Webster et al., 2007, 1, 12, 13–14.

  56. 56.

    Rosenmeier et al., 2002, 183, 185, 188–89.

  57. 57.

    Haug et al., 2003, 1733 (Fig. 2).

  58. 58.

    Lane et al., 2014, 93, 95.

  59. 59.

    Gao et al., 2008; Cole-Dai, 2010, 824–39.

  60. 60.

    Hammer et al., 1980, 235.

  61. 61.

    Herron, 1982.

  62. 62.

    Hammer, 1984, 51–65; Clausen et al., 1997, 26,707–23.

  63. 63.

    Zielinski, 1995, 20,939, 20,944; Clausen et al., 1997, 26,707–23.

  64. 64.

    For instance: Cole-Dai et al., 2000, 24,435, 24,438–39; Kurbatov et al., 2006. On the missing GISP2 section, Zielinski, 1995, 20,940, 20,949, 20,953.

  65. 65.

    Traufetter et al., 2004, 141; Severi et al., 2007, 367–74; Larsen et al., 2008.

  66. 66.

    Baillie, 2008. Recently supported by Sigl et al., 2015, 543.

  67. 67.

    Ferris et al., 2011; Plummer et al., 2012, 1931, 1933–36.

  68. 68.

    Sigl et al., 2013, 1159.

  69. 69.

    Motizuki et al., 2014, 785, 798.

  70. 70.

    Sigl et al., 2014, 693, 694, 695.

  71. 71.

    Sigl et al., 2015, 544, 545, 547–48; also Büntgen et al., 2016.

  72. 72.

    Aizen et al., 2016, Fig. 5a.

  73. 73.

    Stothers and Rampino, 1983b, 6357, 6362–63, 6369; Stothers, 1984, 344–45.

  74. 74.

    Simarski, 1992, 3–5; Kelly and Sear, 1984, 740–43; Bradley, 1988, 221–43; Schmincke, 2004, 259–72.

  75. 75.

    Luterbacher and Pfister, 2015, 246.

  76. 76.

    Hammer et al., 1980, 233, 235. This ‘White River Ash’ eruption was redated recently to 833–50/847 ± 1: Jensen et al., 2014, 875–78.

  77. 77.

    Stothers and Rampino, 1983a, 412, 1983b, 6362; Rampino, 1988, 88. Cf. Heming, 1974, 1259.

  78. 78.

    Stothers, 1999, 717.

  79. 79.

    Traufetter et al., 2004, 141, 145; McKee et al., 2011, 27–37, 2015, 1–7. Stother’s 540 ± 90 date was shown as well to be a mistake.

  80. 80.

    Keys, 2000, 277–78, 86–91.

  81. 81.

    Sigl et al., 2014, 695, 2015, 547–48. The second eruption had been earlier put in the tropics: Ferris et al., 2011 (who dated it to c. 535) proposed a “low latitudes” site and Larsen et al., 2008 (who dated it to 533/4 ± 2) were confident the eruption took place near the equator. Larsen et al., 2008, assigned the first event (with a 529 ± 2 date) a “more northerly source.”

  82. 82.

    Sigl et al., 2013, 2015.

  83. 83.

    Both seem to register only in Greenlandic ice: Sigl et al., 2015, 547 (Fig. 5).

  84. 84.

    Andrea Burke, personal correspondence, June 20, 2016.

  85. 85.

    Tilling et al., 1984, 747–49; Espíndola et al., 2000, 90, 93, 102.

  86. 86.

    Nooren et al., 2009, 97, 101, 106–07. It is not specified why the dendroclimatological data for widespread cooling c. 536 was overlooked. Recently, Nooren et al., 2017 has again assigned the eruption to El Chichón.

  87. 87.

    Dull et al., 2010. Dull had previously dated the eruption to 410–535 and, more precisely, c. 430, Dull et al., 2001, 25, 27; Dull, 2004, 238, 243. A wide mid-fourth- to mid-sixth-century window is advanced independently in Mehringer et al., 2005, 199, 203–04, and Kitamura, 2010, 28.

  88. 88.

    Sigl et al., 2015, Extended Data Tab. 4 puts the Ilopango event at 540.

  89. 89.

    Toohey et al., 2016, 410.

  90. 90.

    Suggested by Larsen et al., 2008, but assigned to 529 ± 2 before being bumped by Baillie to c. 535.

  91. 91.

    Suzuki and Nakada, 2007, 1545, 1565; Soda, 1996, 40.

  92. 92.

    Sigl et al., 2015, 547; Gill Plunkett personal communication June 20 and 22, 2016.

  93. 93.

    Toohey et al., 2016, 406.

  94. 94.

    Gregory of Tours, Marius of Avenches, John of Biclaro, Victor of Tunnuna, and Isidore of Seville mention nothing plausibly related to Mediterranean sun dimming 536–7.

  95. 95.

    Pearson et al., 2012, 3402–14. One might also question why Cassiodorus had to inform his deputy about the dust veil (see above). If it were a major event, would he not have known? See also Grattan and Pyatt, 1999, 173–74, 77–78; Arjava, 2005, 73–94. Not long before the important study of Larsen et al., 2008, which established evidence for a volcanic origin of 536 clouding at both poles, Larsen advised Arjava (p. 77 n. 24) “nothing of interest” was found in Greenlandic ice layers between 531 and 550.

  96. 96.

    Arjava, 2005, 81–83.

  97. 97.

    Rampino, 1988, 87–88.

  98. 98.

    The modeling employed written accounts of clouding duration to help constrain the height of the eruption column. However, ice core data was used to establish the eruption’s latitude, which is the important factor for understanding the latitudinal spread of volcanic aerosols. Matt Toohey, personal correspondence, November 1 and 2, 2016.

  99. 99.

    For instance: cf. Tabs. 1 and 7 in Principe et al., 2004, 705, 716–17, 719.

  100. 100.

    Oppenheimer and Pyle, 2009, 444; Mrgić, 2004, 238. Others, notably Rosi and Santacroce, 1983, 250, consider the most mortal Vesuvius eruption that of 472.

  101. 101.

    Rosi and Santacroce, 1983, 250–51, 253–55; Pearson et al., 2012, 3406 (Fig. 4). On 472: Kostick and Ludlow, 2015, 8–13.

  102. 102.

    Marcellinus Comes, Chronicle, 25; John Malalas, Chronicle, 14.42, 205–06; Chronicle Paschale, 90–91. For discussion, Stothers and Rampino, 1983b, 6361–62; Kostick and Ludlow, 2015, 8–13. These scientists also link Hydatius’ account (35) of poor weather in northern Portugal to this event (Chronicon, ed. T. Mommsen MGH AA XI p. 35), though Hydatius’ text stops in 469 and this passage should be fixed a late 460s date.

  103. 103.

    Procopius, 1919, VI.4, 324–27.

  104. 104.

    Cassiodorus, 1886, 261–62. Discussion: Macfarlane, 2009, 109–11; Cioni et al., 2011, 789–810.

  105. 105.

    See Principe et al., 2004, 705–07, 710 who attribute a 14C dated tephra layer to 450 ± 50 to 536 (not 472 or 512) and speak of “an explosive eruption” that “must have occurred” considering evidence for Mediterranean clouding. Stothers and Rampino note 536 was “probably not” Mediterranean in origin, but Vesuvius may have erupted after Procopius left Campania: 1983b, 6362, 6367.

  106. 106.

    Arrighi et al., 2004.

  107. 107.

    http://climatechange.umaine.edu/colle_gnifetti_2013_.

  108. 108.

    Clube and Napier, 1991, 49; Baillie, 1994, 216, 1999; Rigby et al., 2004, 123–26. Further discussion of the impact of extraterrestrial impactors: Napier, 2014, 391–92.

  109. 109.

    For instance: Stothers, 2002, 4; D’Arrigo et al., 2003, 257.

  110. 110.

    Baillie, 2008.

  111. 111.

    Rigby et al., 2004, 123–26.

  112. 112.

    Abbott et al., 2008.

  113. 113.

    Abbott et al., 2014a, 2014b.

  114. 114.

    Kostick and Ludlow, 2015, 15.

  115. 115.

    Baillie, 1994, 216; Arjava, 2005, 79, 80, 93.

  116. 116.

    See note 92 above.

  117. 117.

    Moorhead, 2001, 147–48, concentrates on Mediterranean mystery clouding, misdates Cassiodorus’ letter to 533, ignores other accounts of sun veiling, and emphasizes the “remarkable ability” of human societies to “bounce back from disasters, including widespread failures of crops.”

  118. 118.

    Tvauri, 2014, 35, is well versed in the paleoclimology of the downturn (30–32) and suggests “primitive” agrarian technology then in Baltic countries made contemporaries especially vulnerable to famines far worse than those of the historical period. He proposes that a “single incident of famine” could erode centuries of population growth.

  119. 119.

    Parry and Carter, 1985.

  120. 120.

    Fischer et al., 2007.

  121. 121.

    The food shortage spectrum: Garnsey, 1988, x, 6, 20–37, 271.

  122. 122.

    McCormick, 2007b, 878–89; cf. Newfield, 2013, 125–48. Later examples: Atwell, 2001, 32, 42–62.

  123. 123.

    The 1257–8 eruption, recently assigned to Samalas, Indonesia, and long known as the largest of the Common Era, did not generate widespread famine. Unlike downturn events, however, dendroclimatology indicates this event did not much affect climate. Stothers, 2000, 361–74; Timmreck et al., 2009; Mann et al., 2012, 202–05; Anchukaitis et al., 2012, 836–37; Esper et al., 2013, 736.

  124. 124.

    For discussion: Stathakopoulos, 2004, 265.

  125. 125.

    Cassiodorus, 1886, 519–20; 12.22 (513–14); 12.27 (521); 12.28 (523–24); 12.26 (520–21).

  126. 126.

    John the Lydian, 1897, 25; Pseudo-Dionysius of Tel-Mahre, Chronicle, 65; Michael the Syrian, Chronique, 220–21.

  127. 127.

    Pseudo-Zachariah Rhetor, Chronicle, 10.1 (399).

  128. 128.

    Procopius documents several tactical siege shortages then: Stathakopoulos, 2004, 270–77.

  129. 129.

    Davis, 2000, 56. Note the reconquest reached Liguria in 538 and this episcopal report was delivered in person in Rome over the winter of 537–38 meaning the dire situation in Liguria is to be assigned to 537. Milan suffered a multi-month-long siege during the war, but in 538–39, also after this report.

  130. 130.

    Charles-Edwards, 2006, 94–95; Williams, 1965, 4. Note the CELT (Corpus of Electronic Texts) transcription of the Annals of Ulster dates the bread failure to 538: www.ucc.ie/celt/published/T100001A/index.html.

  131. 131.

    Weisburd, 1985, 93. Weisburd implies Chang State’s summer snow and autumnal famine occurred in 536 in the text, but the map caption (also p. 93) seems to date these events to 537.

  132. 132.

    Grove and Rackham, 2001, 143–44; Diaz and Trouet, 2014, 168.

  133. 133.

    Toohey et al., 2016, 401, 406, 408–09, 410, Fig. 2.

  134. 134.

    Ge et al., 2010, Figs. 2 and 3.

  135. 135.

    Axboe, 1999, 186–88, 2001, 119–35. Bondesson and Bondesson, 2012, 167–70, discuss a twenty-item deposit dating to the second quarter of the sixth century.

  136. 136.

    Discussed in Gräslund and Price, 2012, 433–34; also Price, 2015, 258–59. Continuity is seen at many settlements.

  137. 137.

    Löwenborg, 2012, 10–13.

  138. 138.

    Gräslund and Price, 2012, 431–36; Löwenborg, 2012, 5–7; Tvauri, 2014, 32–34, 35–39, and references therein. Detailed discussion of a sixth-century site where bread was found as a burial offering: Arrhenius, 2013, 1–14.

  139. 139.

    Löwenborg, 2012, 5, 8–10, 15–17, 19–23; Tvauri, 2014, 39–40, 42–43, 44–47, 48.

  140. 140.

    The Fimbulwinter was recorded first in the late Viking period and long thought by modern scholars to be rooted in the climatic transition away from a warm Scandinavia Bronze Age about 600–450 bce: Pettersson, 1914, 24. More recently it was assigned to the downturn: Axboe, 1999, 187; Gräslund and Price, 2012, 436–40.

  141. 141.

    Gill, 2000, 228–33, 245, 287, 313, 318.

  142. 142.

    Toohey et al., 2016, 401, 406, 408–09, 410, Fig. 2.

  143. 143.

    For example: Koder, 1996, 277; Farquharson, 1996, 266; Houston, 2000, 73, 74; Gräslund and Price, 2012, 433, 438; Löwenborg, 2012, 7, 17–18, 22; Tvauri, 2014, 32, 35, 36, 46, 48.

  144. 144.

    Bondesson and Bondesson, 2014, 61–67.

  145. 145.

    Palynology indicates a sixth- or seventh-century date for the wide sowing of rye in Estonia: Tvauri, 2014, 30, 47–48, 49.

  146. 146.

    Justinianic Plague: Stathakopoulos, 2004, 110–54; Horden, 2005, 134–60; Little, 2007.

  147. 147.

    Cheyette, 2008, 155–56; Gräslund and Price, 2012, 434; Löwenborg, 2012, 7, 17, 19, 24; Tvauri, 2014, 35; Headrick, 2012, 39–40; Kostick and Ludlow, 2015, 16. Long ago, Farquharson emphasized that the downturn was part of “an extraordinary clustering of events,” which included pandemic and epizootic disease: 1996, 267.

  148. 148.

    Maddicott, 1997, 10–11, 17.

  149. 149.

    Campbell has observed that the Black Death’s arrival in England forestalled a sequence of exceptionally poor harvests from creating famine: Campbell, 2010, 301–04; Campbell, 2012, 140, 144–47, 159.

  150. 150.

    Stathakopoulos, 2003, 254 observes that Seibel lumped this Justinianic Plague and mystery clouding together as though they were causally associated in his 1857 work. Recent linkages include: Brown, 2001, 92–94; Stathakopoulos, 2003, 253–54, 2007, 100; McCormick, 2003, 20–21, n.33; Horden, 2005, 152–53; Sallares, 2007, 284–85; McCormick et al., 2012, 198–99; Gräslund and Price, 2012, 433–34; Lee, 2013, 290; Sigl et al., 2015, 548; Haldon et al., 2014, 123; Izdebski et al., 2015.

  151. 151.

    Though low-resolution paleoclimatology now illuminates a pronounced humid period setting in about 550 in Central Africa: Oslisly et al., 2013. In Western and Northern Africa, there is evidence for dry conditions. Low-resolution hydroclimate proxies in Chad and Algeria identify the sixth century as fitting into a two- or three-century dry period. In some proxies from Ghana and Senegal, this dryness is part of much longer-term aridity. In others, from Nigeria and Cameroon, dry conditions appear to set in abruptly in the sixth century. Reconstructions from Eastern Africa are more variable. The sixth century is the last of a long humid period in parts of Kenya. But proxies from other areas, like Tanzania, indicate dry conditions setting in abruptly in the mid-sixth century. Conversely, wetness sets in suddenly in Rwanda, Namibia, and northeast South Africa in the mid-sixth century: Nash et al., 2016, 6–8.

  152. 152.

    Keys, 2000, 16–23.

  153. 153.

    For example: Sallares, 2007, 284–85; Stathakopoulos, 2007, 100; also Lee, 2013, 290. Horden expressed skepticism, Brown thought the temperature sensitivity of plague-bearing rodent fleas problematic to Key’s theory, and McCormick suggested that the connection was more complex than Keys allowed, though he too thought that the two events connected via the effect of climate change on rodent populations: Horden, 2005, 152–53; Brown, 2001, 92–94; M. McCormick, 2003, 20–21, n.33.

  154. 154.

    Biraben and Le Goff, 1975, 50, 58, 64; Sarris, 2002, 169, 170–72; Sallares, 2007, 251, 285–86 thought the plague popped up closer to home, possibly in Egypt.

  155. 155.

    Morelli et al., 2010, 1140–3; Harbeck et al., 2013; Wagner et al., 2014, 323; Feldman et al., 2016.

  156. 156.

    Stathakopoulos, 2003, 254.

  157. 157.

    McCormick, 2003, n.33; McCormick, 2007a, 303–04.

  158. 158.

    McCormick et al., 2012, 198–99.

  159. 159.

    It is not limited to cold climates or seasons, but pneumonic plague does generally require close contact for transmission. Sallares, 2007, 241–42, 286.

  160. 160.

    Unless the disease became endemic or enzootic following the initial introduction. Justinianic recurrences: Biraben and Le Goff, 1975, 58–60; Stathakopoulos, 2004, 113–24; Horden, 2005, 138–39, n.6.

  161. 161.

    Schmid et al., 2015, 3020–25; Kausrud et al., 2010, 112; Ben-Ari et al., 2011. Campbell has demonstrated the Black Death occurred, in Europe, within a distinct climatic anomaly: Campbell, 2010, 287, 300–05; Campbell, 2012, 144–47.

  162. 162.

    McMichael, 2015.

  163. 163.

    Though see Kausrud et al., 2010.

  164. 164.

    Bos et al., 2016; Seifert et al., 2016.

  165. 165.

    The same would apply to other mosquito-borne diseases. In Europe, both vivax and malariae varieties of malaria were well established south and north of the Alps by 550. Gowland and Western, 2012; Newfield, 2017.

  166. 166.

    See note 5 above.

  167. 167.

    For example: Stathakopoulos, 2004, 166–67, 268; Cheyette, 2008, 155–56; Devroey and Jaubert, 2011, 10; Izdebski et al., 2015.

  168. 168.

    Farquharson, 1996, 267.

  169. 169.

    Arrhenius, 2013, 13.

  170. 170.

    Widgren, 2012, 126, 131–33; Nunn, 2007, 9. In Satingpra, Thailand, a downturn drought is seen as spurring major irrigation works: Stargardt, 2014, 129–30.

  171. 171.

    Houston, 2000, 71, 74.

  172. 172.

    Dahlin and Chase, 2014, 127–55.

  173. 173.

    Lucero, 1999, 814–22.

  174. 174.

    Dunning et al., 2012, 3652–57.

  175. 175.

    Turner and Sabloff emphasize spatial and temporal variability in the effects of Maya droughts: Turner and Sabloff, 2012, 13,908–14.

  176. 176.

    A survey of late antique Mediterranean famines: Stathakopoulos, 2004, 23–30, 35–56.

  177. 177.

    Smit and Wandel, 2006, 282–92; Berkes, 1993, 1–10.

  178. 178.

    Löwenborg, 2012, 22–23.

References

  • Abbott, D.H. et al. “Magnetite and Silicate Spherules from the GISP2 Core at the 536 A.D. Horizon.” In American Geophysical Union, Fall Meeting 2008 Abstracts, 2008.

    Google Scholar 

  • Abbott, D.H. et al. “What Caused Terrestrial Dust Loading and Climate Downturns Between A.D. 533 and 540?” Geological Society of America Special Papers 505 (2014a): 421–38.

    Google Scholar 

  • Abbott, D.H. et al. “Calendar-Year Dating of the Greenland Ice Sheet Project 2 (GISP2) Ice Core from the Early Sixth Century Using Historical, Ion, and Particulate Data.” Geological Society of America Special Papers 505 (2014b): 411–20.

    Google Scholar 

  • Aizen, E.M. et al. “Abrupt and Moderate Climate Changes in the Mid-Latitudes of Asia During the Holocene.” Journal of Glaciology 62 (2016): 411–39.

    Google Scholar 

  • Anchukaitis, K.J. et al. “Tree Rings and Volcanic Cooling.” Nature Geoscience 5 (2012): 836–37.

    Google Scholar 

  • Arjava, Antti. “The Mystery Cloud of 536 CE in the Mediterranean Sources.” Dumbarton Oaks Papers 59 (2005): 73–94.

    Google Scholar 

  • Arrhenius, B. “Helgö in the Shadow of the Dust Veil, 536–37.” Journal of Archaeology and International History 5 (2013): 1–14.

    Google Scholar 

  • Arrighi, S. et al. “Recent Eruptive History of Stromboli (Aeolian Islands, Italy) Determined from High-Accuracy Archeomagnetic Dating.” Geophysical Research Letters 31 (2004): L19603.

    Google Scholar 

  • Aston, W.G. Nihongi: Chronicles of Japan from the Earliest Times to A.D. 697. London: Kegan Paul, Trench, Trübner and Co., 1896.

    Google Scholar 

  • Atwell, William. “Volcanism and Short-Term Climatic Change in East Asian and World History c.1200–1699.” Journal of World History 12 (2001): 29–99.

    Google Scholar 

  • Axboe, M. “The Year 536 and the Scandinavian Gold Hoards.” Medieval Archaeology 43 (1999): 186–88.

    Google Scholar 

  • Axboe, M. “Amulet Pendants and a Darkened Sun: On the Function of the Gold Bracteates and a Possible Motivation for the Large Gold Hoards.” In Roman Gold and the Development of the Early Germanic Kingdoms: Aspects of Technical, Socio-Political, Socio-Economic, Artistic and Intellectual Development, A.D. 1–500: Symposium in Stockholm 14–16 November 1997, edited by B. Magnus, 119–36. Stockholm: Almqvist & Wiksell International, 2001.

    Google Scholar 

  • Baillie, M.G.L. “Dendrochronology Raises Questions About the Nature of the AD 536 Dust-Veil Event.” Holocene 4 (1994): 212–17.

    Google Scholar 

  • Baillie, M.G.L. Exodus to Arthur: Catastrophic Encounters with Comets. London: B.T. Batsford, 1999.

    Google Scholar 

  • Baillie, M.G.L. “Proposed Re-Dating of the European Ice Core Chronology by Seven Years Prior to the 7th Century AD.” Geophysical Research Letters 35 (2008): L15813.

    Google Scholar 

  • Ben-Ari, T.S. et al. “Plague and Climate: Scales Matter.” PLoS Pathogens 7 (2011): e100216.

    Google Scholar 

  • Berkes, F. “Traditional Ecological Knowledge in Perspective.” In Traditional Ecological Knowledge: Concepts and Cases, edited by J. Inglis, 1–10. Ottawa: International Development Research Centre, 1993.

    Google Scholar 

  • Biraben, J.N., and J. Le Goff. “The Plague in the Early Middle Ages.” In Biology of Man in History, edited by R. Forster and O.A. Ranum, 48–80. Baltimore, MD: Johns Hopkins University Press, 1975.

    Google Scholar 

  • Bondesson, L., and T. Bondesson. “Barbarisk Imitation Av Bysantinsk Solidus: Ett Soloffer På Själland.” Fornvännen 107 (2012): 167–70.

    Google Scholar 

  • Bondesson, L., and T. Bondesson. “On the Mystery Cloud of AD 536, A Crisis in Dispute and Epidemic Ergotism: A Linking Hypothesis.” Danish Journal of Archaeology 3 (2014): 61–67.

    Google Scholar 

  • Bos, Kirsten I. et al. “Eighteenth Century Yersinia Pestis Genomes Reveal the Long-Term Persistence of an Historical Plague Focus.” eLife 5 (2016): e12994.

    Google Scholar 

  • Bradley, R.S. “The Explosive Volcanic Eruption Signal in Northern Hemisphere Continental Temperature Records.” Climatic Change 12 (1988): 221–43.

    Google Scholar 

  • Briffa, K.R. “Annual Climate Variability in the Holocene: Interpreting the Message of Ancient Trees.” Quaternary Science Reviews 19 (2000): 87–105.

    Google Scholar 

  • Brown, Neville. History and Climate Change: A Eurocentric Perspective. London: Routledge, 2001.

    Google Scholar 

  • Büntgen, Ulf, and W. Tegel. “European Tree-Ring Data and the Medieval Climate Anomaly.” PAGES News 19 (2011): 14–15.

    Google Scholar 

  • Büntgen, Ulf et al. “2500 Years of European Climate Variability and Human Susceptibility.” Science 331 (2011): 578–82.

    Google Scholar 

  • Büntgen, Ulf et al. “Cooling and Societal Change During the Late Antique Little Ice Age from 536 to Around 660 AD.” Nature Geoscience 9 (2016): 231–36.

    Google Scholar 

  • Cameron, Averil. Procopius and the Sixth Century. Berkeley: University of California Press, 1985.

    Google Scholar 

  • Campbell, Bruce M.S. “Nature as Historical Protagonist: Environment and Society in Pre-Industrial England.” The Economic History Review 63 (2010): 281–314.

    Google Scholar 

  • Campbell, Bruce M.S. “Grain Yields on English Demesnes After the Black Death.” In Town and Countryside in the Age of the Black Death: Essays in Honour of John Hatcher, edited by M. Bailey and S. Rigby, 121–74. Turnhout, Belgium: Brepols, 2012.

    Google Scholar 

  • Carleton, W.C. et al. “A Reassessment of the Impact of Drought Cycles on the Classic Maya.” Quaternary Science Reviews 105 (2014): 151–61.

    Google Scholar 

  • Cassiodorus. The Letters of Cassiodorus: Being a Condensed Translation of the Variae Epistolae of Magnus Aurelius Cassiodorus Senator with an Introduction, by Thomas Hodgkin. Translated by Thomas Hodgkin. London: Henry Fowde, 1886.

    Google Scholar 

  • Charles-Edwards, T.M. The Chronicle of Ireland. Liverpool: Liverpool University Press, 2006.

    Google Scholar 

  • Cheyette, Frederic. “The Disappearance of the Ancient Landscape and the Climatic Anomaly of the Early Middle Ages: A Question to Be Pursued.” Early Medieval Europe 16 (2008): 127–65.

    Google Scholar 

  • Christiansen, B., and F. Ljungqvist. “The Extra-Tropical Northern Hemisphere Temperature in the Last Two Millennia: Reconstructions of Low-Frequency Variability.” Climate of the Past 8 (2012): 765–86.

    Google Scholar 

  • Chronicon Paschale 284–628 AD, edited by M. Whitby and M. Whitby. Liverpool: Liverpool University Press, 1989.

    Google Scholar 

  • Chu, Guoqiang et al. “Seasonal Temperature Variability During the Past 1600 Years Recorded in Historical Documents and Varved Lake Sediment Profiles from Northeastern China.” The Holocene 22 (2011): 785–92.

    Google Scholar 

  • Churakova, Olga V. et al. “A Cluster of Stratospheric Volcanic Eruptions in the AD 530s Recorded in Siberian Tree Rings.” Global and Planetary Change 122 (2014): 140–50.

    Google Scholar 

  • Cioni, R. et al. “The 512 AD Eruption of Vesuvius: Complex Dynamics of a Small Scale Subplinian Event.” Bulletin of Volcanology 73 (2011): 789–810.

    Google Scholar 

  • Clausen, H.B. et al. “A Comparison of the Volcanic Records Over the Past 4000 Years from the Greenland Ice Core Project and Dye 3 Cores.” Journal of Geophysical Research 102 (1997): 26707–23.

    Google Scholar 

  • Clube, S., and W. Napier. “Catastrophism Now.” Astronomy Now 5 (1991): 46–49.

    Google Scholar 

  • Cole-Dai, J. “Volcanoes and Climate.” Wiley Interdisciplinary Reviews: Climate Change 1 (2010): 824–39.

    Google Scholar 

  • Cole-Dai, J. et al. “4100 Year Record of Explosive Volcanism from an East Antarctic Ice Core.” Journal of Geophysical Research: Atmospheres 105 (2000): 24431–41.

    Google Scholar 

  • Cook, Edward R. et al. “Millennia-Long Tree-Ring Records from Tasmania and New Zealand: A Basis for Modelling Climate Variability and Forcing, Past, Present and Future.” Journal of Quaternary Science 21 (2006): 689–99.

    Google Scholar 

  • Cook, Edward R. et al. “Old World Megadroughts and Pluvials During the Common Era.” Science Advances 1 (2015): e1500561.

    Google Scholar 

  • Curtis, J. et al. “Climate Variability on the Yucatan Peninsula (Mexico) During the Past 3500 Years, and Implications for Maya Cultural Evolution.” Quaternary Research 46 (1996): 37–47.

    Google Scholar 

  • Dahlin, B.H., and A.F. Chase. “A Tale of Three Cities: Effects of the AD 536 Event in the Lowland Maya Heartland.” In The Great Maya Droughts in Cultural Context: Case Studies in Resilience and Vulnerability, edited by G. Iannone, 127–55. Boulder: University of Colorado Press, 2014.

    Google Scholar 

  • D’Arrigo, Rosanne et al. “Dendroclimatological Evidence for Major Volcanic Events of the Past Two Millennia.” In Volcanism and the Earth’s Atmosphere, edited by Alan Robock and Clive Oppenheimer, 255–61. Washington, DC: American Geophysical Union, 2003.

    Google Scholar 

  • Davis, R. The Book of Pontiffs (Liber Pontificalis): The Ancient Biographies of the First Ninety Roman Bishops to AD 715. Liverpool: Liverpool University Press, 2000.

    Google Scholar 

  • Devroey, J.P., and A.N. Jaubert. “Family, Income and Labour around the North Sea, 500–1000.” In Making a Living: Family, Labour and Income, edited by E. Vanhaute. Turnhout: Brepols, 2011.

    Google Scholar 

  • Diaz, H., and V. Trouet. “Some Perspectives on Societal Impacts of Past Climatic Changes.” History Compass 12 (2014): 160–77.

    Google Scholar 

  • Dull, R. “Lessons from the Mud, Lessons from the Maya: Paleoecological Records of the Tierra Blanca Joven Eruption.” Geological Society of America Special Papers 375 (2004): 237–44.

    Google Scholar 

  • Dull, R. et al. “Volcanism, Ecology and Culture: A Reassessment of the Volcán Ilopango TBJ Eruption in the Southern Maya Realm.” Latin American Antiquity 12 (2001): 25–44.

    Google Scholar 

  • Dull, R. et al. “Did the Ilopango TBJ Eruption Cause the 536 Event.” In AGU Fall Meeting, Abstracts, 2010.

    Google Scholar 

  • Dunning, Nicholas P. et al. “Kax and Kol: Collapse and Resilience in Lowland Maya Civilization.” Proceedings of the National Academy of Sciences 109 (2012): 3652–57.

    Google Scholar 

  • Esper, Jan et al. “European Summer Temperature Response to Annually Dated Volcanic Eruptions Over the Past Nine Centuries.” Bulletin of Volcanology 75 (2013): 1–14.

    Google Scholar 

  • Esper, Jan et al. “Signals and Memory in Tree-Ring Width and Density Data.” Dendrochronologia 35 (2015): 62–70.

    Google Scholar 

  • Espíndola, J.M. et al. “Volcanic History of El Chichón Volcano (Chiapas, Mexico) During the Holocene, and Its Impact on Human Activity.” Bulletin of Volcanology 62 (2000): 90–104.

    Google Scholar 

  • Farquharson, P. “Byzantium, Planet Earth and the Solar System.” In The Sixth Century: End or Beginning, edited by P. Allen and E. Jeffreys, 263–69. Brisbane: Australian Association for Byzantine Studies, 1996.

    Google Scholar 

  • Feldman, M. et al. “A High-Coverage Yersinia pestis Genome from a Sixth-Century Justinianic Plague Victim.” Molecular Biology and Evolution 33 (2016): 2911–23.

    Google Scholar 

  • Ferris, Dave G. et al. “South Pole Ice Core Record of Explosive Volcanic Eruptions in the First and Second Millennia A.D. and Evidence of a Large Eruption in the Tropics around 535 A.D.” Journal of Geophysical Research: Atmospheres 116 (2011): D17308.

    Google Scholar 

  • Fischer, E.M. et al. “European Climate Response to Tropical Volcanic Eruptions over the Last Half Millennium.” Geophysical Research Letters 34 (2007): L05707.

    Google Scholar 

  • Gao, Chaochao et al. “Volcanic Forcing of Climate Over the Past 1500 Years: An Improved Ice Core-Based Index for Climate Models.” Journal of Geophysical Research: Atmospheres 113 (2008): D23111.

    Google Scholar 

  • García-Suárez, A. et al. “Climate Signal in Tree-Ring Chronologies in a Temperate Climate: A Multi-Species Approach.” Dendrochronologia 27 (2009): 183–98.

    Google Scholar 

  • Garnsey, P. Famine and Food Supply in the Graeco-Roman World: Responses to Risk and Crisis. New York: Cambridge University Press, 1988.

    Google Scholar 

  • Ge, Q.S. et al. “Temperature Variation Through 2000 Years in China: An Uncertainty Analysis of Reconstruction and Regional Difference.” Geophysical Research Letters 37 (2010): L03703.

    Google Scholar 

  • Gill, Richardson. The Great Maya Droughts: Water, Life, and Death. Albuquerque: University of New Mexico Press, 2000.

    Google Scholar 

  • Gill, R.B., and J. Keating. “Volcanism and Mesoamerican Archaeology.” Ancient Mesoamerica 13 (2002): 125–40.

    Google Scholar 

  • Gowland, R.L., and A.G. Western. “Morbidity in the Marshes: Using Spatial Epidemiology to Investigate Skeletal Evidence for Malaria in Anglo-Saxon England (AD 410–1050).” American Journal of Physical Anthropology 147 (2012): 301–11.

    Google Scholar 

  • Gräslund, B., and N. Price. “Twilight of the Gods? The ‘Dust Veil Event’ of AD 536 in Critical Perspective.” Antiquity 86 (2012): 428–43.

    Google Scholar 

  • Grattan, J.P., and F.B. Pyatt. “Volcanic Eruptions Dry Fogs and the European Palaeoenvironmental Record: Localised Phenomena or Hemispheric Impacts.” Global and Planetary Change 21 (1999): 173–79.

    Google Scholar 

  • Grove, A.T., and Oliver Rackham. The Nature of Mediterranean Europe: An Ecological History. New Haven, CT: Yale University Press, 2001.

    Google Scholar 

  • Haldon, John et al. “The Climate and Environment of Byzantine Anatolia: Integrating Science, History, and Archaeology.” Journal of Interdisciplinary History 45 (2014): 113–61.

    Google Scholar 

  • Hammer, C.U. “Traces of Icelandic Eruptions in the Greenland Ice Sheet.” Jokull 34 (1984): 51–65.

    Google Scholar 

  • Hammer, C.U. et al. “Greenland Ice Sheet Evidence of Post-Glacial Volcanism and Its Climatic Impact.” Nature 288 (1980): 230–35.

    Google Scholar 

  • Harbeck, M. et al. “Yersinia Pestis DNA from Skeletal Remains from the 6th Century Reveals Insights into Justinianic Plague.” PLoS Pathogens 9 (2013): e1003349.

    Google Scholar 

  • Haug, G. et al. “Climate and the Collapse of Maya Civilization.” Science 299 (2003): 1731–35.

    Google Scholar 

  • Headrick, D. “The Medieval World, 500 to 1500 CE.” In A Companion to Global Environmental History, edited by J.R. McNeill and E.S. Mauldin, 39–56. Hoboken, NJ: Wiley, 2012.

    Google Scholar 

  • Heming, R.F. “Geology and Petrology of Rabaul Caldera, Papua New Guinea.” Geological Society of America Bulletin 85 (1974): 1253–64.

    Google Scholar 

  • Herron, Michael M. “Impurity Sources of F−, Cl−, NO3− and SO42− in Greenland and Antarctic Precipitation.” Journal of Geophysical Research: Oceans 87 (1982): 3052–60.

    Google Scholar 

  • Hodell, David A. et al. “Possible Role of Climate in the Collapse of Classic Maya Civilization.” Nature 375 (1995): 391–94.

    Google Scholar 

  • Hodell, David A. et al. “Solar Forcing of Drought Frequency in the Maya Lowlands.” Science 292 (2001): 1367–70.

    Google Scholar 

  • Hodell, David A. et al. “Terminal Classic Drought in the Northern Maya Lowlands Inferred from Multiple Sediment Cores in Lake Chichancanab (Mexico).” Quaternary Science Reviews 24 (2005): 1413–27.

    Google Scholar 

  • Holzhauser, H. et al. “Glacier and Lake-Level Variations in West-Central Europe Over the Last 3500 Years.” The Holocene 15 (2005): 789–801.

    Google Scholar 

  • Horden, P. “Mediterranean Plague in the Age of Justinian.” In The Cambridge Companion to the Age of Justinian, edited by M. Mass, 134–60. New York: Cambridge University Press, 2005.

    Google Scholar 

  • Houston, Margaret S. “Chinese Climate, History, and Stability in A.D. 536.” In The Years Without Summer: Tracing A.D. 536 and Its Aftermath, edited by J. Gunn, 71–77. Oxford: Archaeopress, 2000.

    Google Scholar 

  • Iles, C. et al. “The Effect of Volcanic Eruptions on Global Precipitation.” Journal of Geophysical Research Letters: Atmospheres 118 (2013): 8770–86.

    Google Scholar 

  • Izdebski, A. et al. “The Environmental, Archaeological and Historical Evidence for Regional Climatic Changes and Their Societal Impacts in the Eastern Mediterranean in Late Antiquity.” Quaternary Science Reviews 136 (2015): 189–208.

    Google Scholar 

  • Jensen, Britta et al. “Transatlantic Distribution of the Alaskan White River Ash.” Geology 42 (2014): 875–78.

    Google Scholar 

  • Jones, P. et al. “High-Resolution Palaeoclimatology of the Last Millennium: A Review of Current Status and Future Prospects.” The Holocene 19 (2009): 3–49.

    Google Scholar 

  • Kausrud, Kyrre L. et al. “Modeling the Epidemiological History of Plague in Central Asia: Palaeoclimatic Forcing on a Disease System over the Past Millennium.” BMC Biology 8 (2010): 112.

    Google Scholar 

  • Kelly, P.M., and C.B. Sear. “Climatic Impact of Explosive Volcanic Eruptions.” Nature 311 (1984): 740–43.

    Google Scholar 

  • Keys, David. Catastrophe: An Investigation into the Origins of the Modern World. New York: Ballantine, 2000.

    Google Scholar 

  • Kitamura, Shigeru. “Two AMS Radiocarbon Dates for the TBJ Tephra from Ilopango Caldera, El Salvador, Central America.” Bulletin of the Faculty of Social Work, Hirosaki Gakuin University 10 (2010): 24–28.

    Google Scholar 

  • Kobashi, Takuro et al. “High Variability of Greenland Surface Temperature Over the Past 4000 Years Estimated from Trapped Air in an Ice Core.” Geophysical Research Letters 38 (2011): L21501.

    Google Scholar 

  • Koder, Johannes. “Climatic Change in the Fifth and Sixth Centuries?” In The Sixth Century – End or Beginning, edited by P. Allen and E.M. Jeffreys, 270–85. Brisbane: Australian Association for Byzantine Studies, 1996.

    Google Scholar 

  • Kostick, C., and F. Ludlow. “The Dating of Volcanic Events and Their Impact Upon European Society, 400–800 CE.” Post Classical Archaeologies 5 (2015): 7–30.

    Google Scholar 

  • Kurbatov, A.V. et al. “A 12,000 Year Record of Explosive Volcanism in the Siple Dome Ice Core, West Antarctica.” Journal of Geophysical Research: Atmospheres 111 (2006): D12307.

    Google Scholar 

  • LaMarche, V., and K. Hirschboeck. “Frost Rings in Trees as Records of Major Volcanic Eruptions.” Nature 307 (1984): 121–26.

    Google Scholar 

  • Lane, C.S. et al. “Beyond the Mayan Lowlands: Impacts of the Terminal Classic Drought in the Caribbean Antilles.” Quaternary Science Reviews 86 (2014): 89–98.

    Google Scholar 

  • Larsen, L.B. et al. “New Ice Core Evidence for a Volcanic Cause of the A.D. 536 Dust Veil.” Geophysical Research Letters 35 (2008): L04708.

    Google Scholar 

  • Lee, A.D. From Rome to Byzantium AD 363 to 565: The Transformation of Ancient Rome. Edinburgh: Edinburgh University Press, 2013.

    Google Scholar 

  • Little, L., ed. Plague and the End of Antiquity: The Pandemic of 541–750. New York: Cambridge University Press, 2007.

    Google Scholar 

  • Löwenborg, D. “An Iron Age Shock Doctrine – Did the AD 536-7 Event Trigger Large-Scale Social Changes in the Mälaren Valley Area?” Journal of Archaeology and International History 4 (2012): 1–29.

    Google Scholar 

  • Lucero, L. “The Collapse of the Classic Maya: A Case for the Role of Water Control.” Archeological Papers of the American Anthropological Association 9 (1999): 35–49.

    Google Scholar 

  • Luterbacher, Jürg, and Christian Pfister. “The Year Without a Summer.” Nature Geoscience 8 (2015): 246–48.

    Google Scholar 

  • Luterbacher, Jürg et al. “A Review of 2000 Years of Paleoclimatic Evidence in the Mediterranean.” In The Climate of the Mediterranean Region: From the Past to the Future, edited by P. Lionello, 87–185. Burlington: Elsevier Science, 2012.

    Google Scholar 

  • Luterbacher, Jürg et al. “European Summer Temperatures Since Roman Times.” Environmental Research Letters 11 (2016): 024001.

    Google Scholar 

  • Lydian, John. Liber de Ostentis et Calendaria Graeca Omnia. Edited by C. Wachsmuth. Leipzig: G. Teubneri, 1897.

    Google Scholar 

  • Macfarlane, R.T. “Vesuvian Narratives: Collisions and Collusions of Man and Volcano.” In Apolline Project 1: Studies on Vesuvius’ North Slope and the Bay of Naples, edited by G. De Simone and R.T. Macfarlane, 103–21. Naples: Università degli studi Suor Orsola Benincasa, 2009.

    Google Scholar 

  • Maddicott, J. “Plague in Seventh-Century England.” Past and Present 156 (1997): 7–54.

    Google Scholar 

  • Malalas, John. The Chronicle of John Malalas. Translated by Elizabeth Jeffreys, Michael Jeffreys, Roger Scott, and Brian Croke. Melbourne: Australian Association for Byzantine Studies, 1986.

    Google Scholar 

  • Mann, M.E. et al. “Underestimation of Volcanic Cooling in Tree-Ring Based Reconstructions of Hemispheric Temperatures.” Nature Geoscience 5 (2012): 202–05.

    Google Scholar 

  • Marcellinus Comes. The Chronicle of Marcellinus: A Translation and Commentary. Translated by Brian Croke. Sydney: Australian Association for Byzantine Studies, 1995.

    Google Scholar 

  • Masson-Delmotte, V. et al. “Information from Paleoclimate Archives.” In Climate Change 2013: The Physical Science Basis: Working Group I Contribution to the Fifth Assessment Report of the Intergovernmental Panel on Climate Change, 383–464. New York: Cambridge University Press, 2014.

    Google Scholar 

  • McCormick, Michael. “Rats, Communications, and Plague: Toward an Ancient and Medieval Ecological History.” Journal of Interdisciplinary History 34 (2003): 1–25.

    Google Scholar 

  • McCormick, Michael. “Toward a Molecular History of the Justinianic Pandemic.” In Plague and the End of Antiquity: The Pandemic of 541–750, edited by L. Little, 290–312. New York: Cambridge University Press, 2007a.

    Google Scholar 

  • McCormick, Michael. “Volcanoes and the Climate Forcing of Carolignian Europe, A.D. 750–950.” Speculum 82 (2007b): 865–96.

    Google Scholar 

  • McCormick, Michael. “What Climate Science, Ausonius, Nile Floods, Rye Farming, and Thatched Roofs Tell Us about the Environmental History of the Roman Empire.” In The Ancient Mediterranean Environment between Science and History, edited by William V. Harris, 61–88. Leiden; Boston: Brill, 2013.

    Google Scholar 

  • McCormick, Michael et al. “Climate Change During and After the Roman Empire: Reconstructing the Past from Scientific and Historical Evidence.” Journal of Interdisciplinary History 43 (2012): 169–220.

    Google Scholar 

  • McKee, C.O. et al. “A Remarkable Pulse of Large-Scale Volcanism on New Britain Island, Papua New Guinea.” Bulletin of Volcanology 73 (2011): 27–37.

    Google Scholar 

  • McKee, C.O. et al. “A Revised Age of AD 667–699 for the Latest Major Eruption at Rabaul.” Bulletin of Volcanology 77 (2015): 65.

    Google Scholar 

  • McMichael, A. “Extreme Weather Events and Infectious Disease Outbreaks.” Virulence 6 (2015): 543–47.

    Google Scholar 

  • Medina-Elizalde, M. et al. “High Resolution Stalagmite Climate Record from the Yucatán Peninsula Spanning the Maya Terminal Classic Period.” Earth and Science Planetary Letters 298 (2010): 255–62.

    Google Scholar 

  • Mehringer, P. et al. “Age and Extent of the Ilopango TBJ Tephra Inferred from a Holocene Chronostratigraphic Reference Section, Lago de Yojoa, Honduras.” Quaternary Research 63 (2005): 199–205.

    Google Scholar 

  • Michael the Syrian. Chronique de Michel Le Syrien Patriarche Jacobite d’Antioche: 1166–1199. Translated by J.B. Chabot, 1901.

    Google Scholar 

  • Moorhead, John. The Roman Empire Divided, 400–700. Harlow: Longman, 2001.

    Google Scholar 

  • Morelli, G. et al. “Yersinia Pestis Genome Sequencing Identifies Patterns of Global Phylogenetic Diversity.” Nature Genetics 42 (2010): 1140–43.

    Google Scholar 

  • Moreno, P.I. et al. “Southern Annular Mode-Like Changes in Southwestern Patagonia at Centennial Timescales Over the Last Three Millennia.” Nature Communications 5 (2014): 4735.

    Google Scholar 

  • Motizuki, Y. et al. “Dating of a Dome Fuji (Antarctica) Shallow Ice Core by Volcanic Signal Synchronization with B32 and EDML1 Chronologies.” Cryosphere Discussions 8 (2014): 769–804.

    Google Scholar 

  • Mrgić, J. “Ash Fell from the Skies to the Earth: The Eruption of the Vesuvius in 1631 AD and the Balkan Lands.” Balcanica 35 (2004): 223–38.

    Google Scholar 

  • Napier, W. “The Role of Giant Comets in Mass Extinctions.” Geological Society of America Special Papers 505 (2014).

    Google Scholar 

  • Nash, David et al. “African Hydroclimatic Variability During the Last 2000 Years.” Quaternary Science Reviews 154 (2016): 1–22.

    Google Scholar 

  • Newfield, Timothy P. “The Causation, Contours and Frequency of Food Shortages in Carolingian Europe, c.750–c.950.” In Crisis en la Edad Media: Modelos, Explicaciones y Representaciones, edited by Pere Benito i Monclús, 117–72. Barcelona: Editorial Milenio Lleida, 2013.

    Google Scholar 

  • Newfield, Timothy P. “Malaria and Malaria-Like Disease in the Early Middle Ages.” Early Medieval Europe 25 (2017): 251–300.

    Google Scholar 

  • Nooren, C.A.M. et al. “Tephrochronological Evidence for the Late Holocene Eruption History of El Chichón, Mexico.” Geofisica Internacional 48 (2009): 97–112.

    Google Scholar 

  • Nooren, Kees et al. “Explosive Eruption of El Chichón Volcano (Mexico) Disrupted 6th Century Maya Civilization and Contributed to Global Cooling.” Geology 45 (2017): 175–78.

    Google Scholar 

  • Nunn, Patrick D. Climate, Environment and Society in the Pacific during the Last Millennium. Amsterdam: Elsevier, 2007.

    Google Scholar 

  • Oppenheimer, C., and D. Pyle. “Volcanoes.” In Physical Geography of the Mediterranean, edited by J. Woodward. New York: Oxford, 2009.

    Google Scholar 

  • Oslisly, Richard et al. “Climatic and Cultural Changes in the West Congo Basin Forests Over the Past 5000 Years.” Phil. Transactions of the Royal Society B 368 (2013): 20120304.

    Google Scholar 

  • Pages 2k Consortium. “Continental-Scale Temperature Variability During the Past Two Millennia.” Nature Geoscience 6 (2013): 339–46.

    Google Scholar 

  • Parker, D.E. “Frost Rings in Trees and Volcanic Eruptions.” Nature 313 (1985): 160–61.

    Google Scholar 

  • Parry, M.L., and T.R. Carter. “The Effect of Climatic Variations on Agricultural Risk.” Climatic Change 7 (1985): 95–110.

    Google Scholar 

  • Pearson, C. et al. “Dendroarchaeology of the Mid-First Millennium AD in Constantinople.” Journal of Archaeological Science 39 (2012): 3402–14.

    Google Scholar 

  • Pettersson, O. Climatic Variation in Historic and Prehistoric Time. Göteborg: W. Zachrissons, 1914.

    Google Scholar 

  • Plummer, C.T. et al. “An Independently Dated 2000-Yr Record from Law Dome, East Antarctica, Including a New Perspective on the 1450s CE Eruption of Kuwae, Vanuatu.” Climate of the Past 8 (2012): 1929–40.

    Google Scholar 

  • Power, Timothy. The Red Sea from Byzantium to the Caliphate: AD 500–1000. Cairo: American University in Cairo, 2012.

    Google Scholar 

  • Price, T.D. Ancient Scandinavia: An Archaeological History from the First Humans to the Vikings. New York: Oxford University Press, 2015.

    Google Scholar 

  • Principe, C. et al. “Chronology of Vesuvius’ Activity from A.D. 79 to 1631 Based on Archaeomagnetism of Lavas and Historical Sources.” Bulletin of Volcanology 66 (2004): 703–24.

    Google Scholar 

  • Procopius. History of the Wars II. Translated by H.B. Dewing. Cambridge, MA: Harvard University Press, 1916.

    Google Scholar 

  • Procopius. History of the Wars III. Translated by H.B. Dewing. Cambridge, MA: Harvard University Press, 1919.

    Google Scholar 

  • Pseudo-Dionysius of Tel-Mahre. Translated by W. Witakowski. Liverpool: Senate House, 1996.

    Google Scholar 

  • Pseudo-Zachariah Rhetor. Chronicle. Translated by R. Phenix and B. C. Horn. Liverpool University Press, 2011.

    Google Scholar 

  • Raible, C. et al. “Tambora 1815 as a Test Case for High Impact Volcanic Eruptions: Earth System Effects.” WIRES Climate Change 7 (2016): 569–89.

    Google Scholar 

  • Rampino, M. “Volcanic Winters.” Annual Review of Earth and Planetary Sciences 16 (1988): 73–99.

    Google Scholar 

  • Rigby, Emma et al. “A Comet Impact in AD 536?” Astronomy and Geophysics 45 (2004): 1.23–1.26.

    Google Scholar 

  • Rosenmeier, M. et al. “A 4000-Year Lacustrine Record of Environmental Change in the Southern Maya Lowlands, Petén, Guatemala.” Quaternary Research 57 (2002): 183–90.

    Google Scholar 

  • Rosi, M., and R. Santacroce. “The A.D. 472 ‘Pollena’ Eruption: Volcanological and Petrological Data for This Poorly-Known, Plinian-Type at Vesuvius.” Journal of Volcanology and Geothermal Research 17 (1983): 249–71.

    Google Scholar 

  • Sallares, R. “Ecology, Evolution, and Epidemiology of Plague.” In Plague and the End of Antiquity: The Pandemic of 541–750, edited by L. Little, 231–89. New York: Cambridge University Press, 2007.

    Google Scholar 

  • Sarris, Peter. “The Justinianic Plague: Origins and Effects.” Continuity and Change 17 (2002): 169–82.

    Google Scholar 

  • Schmid, Boris V. et al. “Climate-Driven Introduction of the Black Death and Successive Plague Reintroductions into Europe.” Proceedings of the National Academy of Sciences 112 (2015): 3020–25.

    Google Scholar 

  • Schmincke, H.-U. “Volcanoes and Climate.” In Volcanism, edited by H.-U. Schmincke, 259–72. New York: Springer, 2004.

    Google Scholar 

  • Schneider, P. “The So-Called Confusion Between India and Ethiopia: The Eastern and Southern Edges of the Inhabited World from the Greco-Roman Perspective.” In Brill’s Companion to Ancient Geography: The Inhabited World in Greek and Roman Tradition, edited by S. Bianchetti, M. Cataudella, and H.J. Gehrke, 184–202. Boston: Brill, 2015.

    Google Scholar 

  • Seifert, Lisa et al. “Genotyping Yersinia Pestis in Historical Plague: Evidence for Long-Term Persistence of Y. Pestis in Europe from the 14th to the 17th Century.” PLoS ONE 11 (2016): e0145194.

    Google Scholar 

  • Severi, M. et al. “Synchronisation of the EDML and EDC Ice Cores for the Last 52 Kyr by Volcanic Signature Matching.” Climate of the Past 3 (2007): 367–74.

    Google Scholar 

  • Sigl, Michael et al. “A New Bipolar Ice Core Record of Volcanism from WAIS Divide and NEEM and Implications for Climate Forcing of the Last 2000 Years.” Journal of Geophysical Research: Atmospheres 118 (2013): 1151–69.

    Google Scholar 

  • Sigl, Michael et al. “Insights from Antarctica on Volcanic Forcing During the Common Era.” Nature Climate Change 4 (2014): 693–97.

    Google Scholar 

  • Sigl, Michael et al. “Timing and Climate Forcing of Volcanic Eruptions for the Past 2500 Years.” Nature 523 (2015): 543–49.

    Google Scholar 

  • Simarski, L. Volcanism and Climate Change. Washington, DC: American Geophysical Union Special Report, 1992.

    Google Scholar 

  • Smit, B., and J. Wandel. “Adaptation, Adaptive Capacity and Vulnerability.” Global Environmental Change 16 (2006): 282–92.

    Google Scholar 

  • Soda, Tstutomu. “Explosive Activities of Haruna Volcano and Their Impacts on Human Life in the Sixth Century AD.” Geographical Reports of Tokyo Metropolitan University 31 (1996): 37–52.

    Google Scholar 

  • Stargardt, J. “Irrigation in South Thailand as a Coping Strategy Against Climate Change: Past and Present.” In Environmental and Climate Change in South and Southeast Asia: How Are Local Cultures Coping?, edited by B. Schuler, 105–37. Leiden: Brill, 2014.

    Google Scholar 

  • Stathakopoulos, Dionysios. “Reconstructing the Climate of the Byzantine World: State of the Problem and Case Studies.” In People and Nature in Historical Perspective, edited by Péter Szábo and József Laszlovszky, 247–61. Budapest: Central European University, 2003.

    Google Scholar 

  • Stathakopoulos, Dionysios. Famine and Pestilence in the Late Roman and Early Byzantine Empire. Burlington, VT: Ashgate, 2004.

    Google Scholar 

  • Stathakopoulos, Dionysios. “Crime and Punishment: The Plague in the Byzantine Empire, 541–750.” In Plague and the End of Antiquity, edited by Lester K. Little, 99–118. New York: Cambridge University Press, 2007.

    Google Scholar 

  • Steig E. et al. “Recent Climate and Ice-Sheet Changes in West Antarctica Compared with Past 2000 Years.” Nature Geoscience 6 (2013): 372–75.

    Google Scholar 

  • Stothers, Richard B. “Mystery Cloud of AD 536.” Nature 307 (1984): 344–45

    Google Scholar 

  • Stothers, Richard B. “Volcanic Dry Fogs, Climate Cooling, and Plague Pandemics in Europe and the Middle East.” Climatic Change 42 (1999): 713–23.

    Google Scholar 

  • Stothers, Richard B. “Climatic and Demographic Consequences of the Massive Volcanic Eruption of 1258.” Climatic Change 45 (2000): 361–64.

    Google Scholar 

  • Stothers, Richard B. “Cloudy and Clear Stratospheres Before A.D.1000 Inferred from Written Sources.” Journal of Geophysical Research: Atmospheres 107 (2002): D23.

    Google Scholar 

  • Stothers, Richard B., and M.R. Rampino. “Historic Volcanism, European Dry Fogs, and Greenland Acid Precipitation, 1500 B.C. to A.D. 1500.” Science 222 (1983a): 411–13.

    Google Scholar 

  • Stothers, Richard B., and M.R. Rampino “Volcanic Eruptions in the Mediterranean Before A.D. 630 From Written and Archaeological Sources.” Journal of Geophysical Research 88 (1983b): 6357–71.

    Google Scholar 

  • Suzuki, Y., and S. Nakada. “Remobilization of Highly Crystalline Felsic Magma by Injection of Mafic Magma: Constraints from the Middle Sixth Century Eruption at Haruna Volcano, Honshu, Japan.” Journal of Petrology 48 (2007): 1543–67.

    Google Scholar 

  • Tan, M. et al. “Cyclic Rapid Warming on Centennial-Scale Revealed by a 2650-Year Stalagmite Record of Warm Season Temperature.” Geophysical Research Letters 30 (2003): 1617.

    Google Scholar 

  • The Koguryo Annals of the Samguk Sagi. Translated by E. Shultz. Seongnam-si, Korea: Academy of Korean Studies Press, 2011.

    Google Scholar 

  • The Silla Annals of the Samguk Sagi. Translated by E. Shultz. Seongnam-si, Korea: Academy of Korean Studies Press, 2012.

    Google Scholar 

  • Thompson, L. et al. “A 1500-Year Record of Tropical Precipitation in Ice Cores from the Quelccaya Ice Cap, Peru.” Science 229 (1985): 971–73.

    Google Scholar 

  • Thompson, L. et al. “Glacial Records of Global Climate: A 1500-Year Tropical Ice Core Record of Climate.” Human Ecology 22 (1994): 83–95.

    Google Scholar 

  • Tilling, R.I. et al. “Holocene Eruptive Activity of El Chichón, Chiapas, Mexico.” Science 224 (1984): 747–50.

    Google Scholar 

  • Timmreck, C. et al. “Limited Temperature Response to the Very Large AD 1258 Volcanic Eruption.” Geophysical Research Letters 36 (2009): L21708.

    Google Scholar 

  • Toohey, M. et al. “Climatic and Societal Impacts of a Volcanic Double Event at the Dawn of the Middle Ages.” Climatic Change 136 (2016): 401–12.

    Google Scholar 

  • Traufetter, F. et al. “Spatio-Temporal Variability in Volcanic Sulphate Deposition Over the Past 2 kyr in Snow Pits and Fir Cones from Amundsenisen, Antarctica.” Journal of Glaciology 50 (2004): 137–46.

    Google Scholar 

  • Turner, B.L., and Jeremy A. Sabloff. “Classic Period Collapse of the Central Maya Lowlands: Insights About Human–Environment Relationships for Sustainability.” Proceedings of the National Academy of Sciences 109 (2012): 13908–14.

    Google Scholar 

  • Tvauri, A. “The Impact of the Climate Catastrophe of 536–537 AD in Estonia and Neighbouring Areas.” Estonian Journal of Archaeology 18 (2014): 30–56.

    Google Scholar 

  • Van Bellen, Simon et al. “Late-Holocene Climate Dynamics Recorded in the Peat Bogs of Tierra Del Fuego, South America.” The Holocene 26 (2015): 489–501.

    Google Scholar 

  • Wagner, D.M. et al. “Yersinia Pestis and the Plague of Justinian 541–543 AD: A Genomic Analysis.” The Lancet 14 (2014): 319–26.

    Google Scholar 

  • Wahl, David et al. “An 8700 Year Paleoclimate Reconstruction from the Southern Maya Lowlands.” Quaternary Science Reviews 103 (2014): 19–25.

    Google Scholar 

  • Webster, J.W. et al. “Stalagmite Evidence from Belize Indicating Significant Droughts at the Time of the Preclassic Abandonment, the Maya Hiatus, and the Classic Maya Collapse.” Palaeogeography, Palaeoclimatology, Palaeoecology 250 (2007): 1–17.

    Google Scholar 

  • Weisburd, Stefi. “Excavating Words: A Geological Tool.” Science News 127 (1985): 91–94.

    Google Scholar 

  • Wickham, Chris. Framing the Early Middle Ages: Europe and the Mediterranean 400–800. New York: Oxford University Press, 2005.

    Google Scholar 

  • Widgren, M. “Climate and Causation in the Swedish Iron Age: Learning from the Present to Understand the Past.” Danish Journal of Geography 112 (2012): 126–34.

    Google Scholar 

  • Williams, J., ed. Annales Cambriae. Wiesbaden: Kraus Reprint Ltd., 1965.

    Google Scholar 

  • Zhang, P. et al. “A Test of Climate, Sun, and Culture Relationships from an 1810-Year Chinese Cave Record.” Science 322 (2008): 940–42.

    Google Scholar 

  • Zielinski, G.A. “Stratospheric Loading and Optimal Depth Estimates of Explosive Volcanism Over the Last 2100 Years Derived from the Greenland Ice Sheet Project 2 Ice Core.” Journal of Geophysical Research: Atmospheres 100 (1995): 20937–55.

    Google Scholar 

Download references

Author information

Authors and Affiliations

Authors

Editor information

Editors and Affiliations

Copyright information

© 2018 The Author(s)

About this chapter

Check for updates. Verify currency and authenticity via CrossMark

Cite this chapter

Newfield, T.P. (2018). The Climate Downturn of 536–50. In: White, S., Pfister, C., Mauelshagen, F. (eds) The Palgrave Handbook of Climate History. Palgrave Macmillan, London. https://doi.org/10.1057/978-1-137-43020-5_32

Download citation

  • DOI: https://doi.org/10.1057/978-1-137-43020-5_32

  • Published:

  • Publisher Name: Palgrave Macmillan, London

  • Print ISBN: 978-1-137-43019-9

  • Online ISBN: 978-1-137-43020-5

  • eBook Packages: HistoryHistory (R0)

Publish with us

Policies and ethics