Earthquakes in Political, Economic, and Cultural History (2024)

Introduction

Since prehistory, human societies have cohabited with seismicity in a “fatal attraction” (the evocative phrase of geophysicist James Jackson), because the advantages of living with earthquakes easily outweigh the disadvantages (Jackson, 2006, p. 1911). This was true of the ancient civilizations in China, Greece, and Palestine, as recorded, for example, in the Bible’s frequent references to earthquakes, and it is evidently true of the modern world. More than half of the world’s largest cities—as many as 60 cities—lie on plate-tectonic boundaries, such as the San Andreas Fault of California, in areas of major seismic activity. They include Ankara, Athens, Beijing, Cairo, Caracas, Delhi, Hong Kong, Istanbul, Jakarta, Karachi, Kathmandu, Lisbon, Lima, Los Angeles, Manila, Mexico City, Naples, Osaka, Rome, San Francisco, Santiago, Shanghai, Singapore, Taipei, Teheran, and Tokyo. Some of them—notably Caracas, Kathmandu, Lisbon, Lima, Los Angeles, Manila, Mexico City, Naples, San Francisco, Teheran, and Tokyo—have suffered major destruction from earthquakes during the past two or three centuries.

In the 19th century, as the study of earthquakes slowly became the science of seismology, Charles Darwin—the great English naturalist who originally made his name as a geologist—experienced a severe earthquake in South America on the coast of Chile, while circumnavigating the planet on HMS Beagle. In his classic travel diary, Voyage of the Beagle, Darwin ranked the earthquake and its impact as the most “deeply interesting” sight of his entire 5-year journey (Darwin, 1989, p. 235). For Darwin, born and brought up in geologically stable England, this was his first personal encounter with the earth’s instability. As he beheld the newly devastated Chilean city of Concepción in 1835, Darwin brooded pessimistically that:

Earthquakes alone are sufficient to destroy the prosperity of any country. If, for instance, beneath England, the now inert subterranean forces should exert those powers which most assuredly in former geological ages they have exerted, how completely would the entire condition of the country be changed! What would become of the lofty houses, thickly packed cities, great manufacturies, the beautiful public and private edifices? If the new period of disturbance were first to commence by some great earthquake in the dead of night, how terrific would be the carnage! England would at once be bankrupt; all papers, records, and accounts would from that moment be lost. Government being unable to collect the taxes, and failing to maintain its authority, the hand of violence and rapine would go uncontrolled. In every large town famine would be proclaimed, pestilence and death following in its train.

(Darwin, 1989, p. 232)

England has never been put through Darwin’s seismic ordeal (although it has been far from earthquake-free, as catalogued in Charles Davison’s A History of British Earthquakes). So it is not possible to verify Darwin’s doom-laden prediction. However, on the basis of reliable seismic records, beginning in China in 780 bc, Darwin was probably wrong to suggest that mid-19th-century England would have struggled to rebound from such a hypothetical geological assault, given its strong government and unparalleled industrial and financial resources, not to speak of its extensive colonial empire. For similar reasons, 20th-century San Francisco and Tokyo rebounded relatively fast from devastating earthquakes, in 1906 and 1923, respectively. In the 21st century, compare the effects of two major earthquakes in 2010, which were less powerful than those in San Francisco and Tokyo but still highly destructive. One of them, of magnitude 7.1, struck New Zealand in the Pacific Ocean, 40 km from the city of Christchurch, yet caused not a single fatality. The other, of magnitude 7.0, struck Haiti in the Caribbean Sea, 25 km from its capital Port-au-Prince, and caused somewhere between 85,000 and 316,000 deaths; the higher figure is the Haitian government’s estimate, which is disputed by international aid agencies. The most significant reason for the huge difference in the fatalities in Christchurch and Port-au-Prince was the reinforced construction of buildings in New Zealand, as compared with the unreinforced construction of buildings in Haiti—a fact that of course depends on the very different degree of political, economic, technological, and scientific development in New Zealand versus Haiti.

As of 2017, the United States has the world’s largest economy, China the second largest, and Japan the third largest. China and Japan are among the world’s most seismically active countries, however seismicity is measured (i.e., according to contemporary seismographic monitoring, the historical earthquake record, or the number of earthquake fatalities). Approximately 22% of the world’s earthquakes of magnitude 6.0 or greater occur in Japan.

On this evidence, one might even argue that destructive earthquakes, for all their horrors, can enhance economic growth over the longer term. “Earthquakes create a lot of business,” remarked the so-called “father of seismology,” John Milne, in the first decade of the 20th century, on the basis of his personal experience of earthquakes in Japan (Herbert-Gustar & Nott, 1980, p. 133). “It is conventional, and by no means inappropriate, to think of disasters in strictly negative terms, but calamities have also often presented opportunities,” wrote historian Kevin Rozario in The Culture of Calamity: Disaster and the Making of Modern America—as happened in San Francisco after the 1906 earthquake. “Americans, especially those in positions of power and influence, have often viewed disasters as sources of moral, political and economic renewal” (Rozario, 2007, p. 3). Indeed, natural disasters can be powerful promoters of corporate and free-market interests, argues social activist Naomi Klein in The Shock Doctrine: The Rise of Disaster Capitalism, with specific reference to the Indian Ocean tsunami produced by an Indonesian earthquake in 2004. Some thinkers have gone so far as to see earthquakes as blessings in disguise. After the seismic destruction of Lisbon, the capital of Portugal, in 1755, the philosopher Immanuel Kant claimed: “Just as we complain of ill-timed or excessive rain, forgetting that rain feeds the springs necessary in our economy, so we denounce earthquakes, refusing to consider whether they too may not bring us good things” (Kendrick, 1956, p. 132). In 1848, in his Principles of Political Economy, economist John Stuart Mill predicted long-term benefits from such disasters, because they obliterated old stock and encouraged manufacturers to introduce efficiency savings in production processes. From a religious and political perspective, Indian political leader Mahatma Gandhi maintained that a great earthquake in north India and Nepal in 1934 was a warning to caste Hindus against the sin of Untouchability. Gandhi publicly proclaimed, “Visitations like droughts, floods, earthquakes and the like, though they seem to have only physical origins are, for me, somehow connected with men’s morals” (Gandhi, 1958, p. 165).

In order to illustrate the claims of the preceding paragraph, eleven earthquakes—out of several hundred historically influential earthquakes—have been selected for their evidently long-term political, economic, and cultural consequences. For example, in politics, a Greek earthquake that damaged Sparta ca. 464 bc was an important factor in starting the crucial Peloponnesian wars between Sparta and Athens in 460 bc, and a Venezuelan earthquake in Caracas in 1812 precipitated Simon Bolívar’s liberation of much of South America from Spanish colonial rule in the 1820s. In economics, a Chinese earthquake in Tangshan in 1976 that coincided with the death of Mao Zedong catalyzed the end of the Cultural Revolution and the rise of China at the turn of the second millennium, and an Indian earthquake in 2001 led to striking industrial growth in the devastated western state of Gujarat that enabled the election of Gujarat’s chief minister, Narendra Modi, as prime minister of India in 2014. In culture, British earthquakes in 1750 initiated the scientific study of seismicity by the fellows of the Royal Society, which undermined the power of the Anglican Church, and the Lisbon earthquake of 1755 was pivotal in freeing Enlightenment science from Catholic religious orthodoxy, as epitomized by Voltaire’s satirical novel Candide, written in response to the earthquake.

The long-term impact of a great earthquake depends on its epicenter, magnitude, and timing—and also on human factors: the political, economic, social, intellectual, religious, and cultural resources specific to a region’s history. Each earthquake-struck society offers its own particular lesson, and yet, taken together, such earth-shattering events have important shared consequences for political, economic, and cultural development on a global scale.

Ancient Greece and Rome

Both Greece and Italy are zones of high seismicity. However, records of Greek and Italian earthquakes do not begin until the 5th century bc. The earthquake in about 464 bc that struck Sparta, in the Peloponnese, is one of the very few earthquakes of classical antiquity that was fairly reliably reported.

Thucydides, writing his History of the Peloponnesian War in the same century, noted that this earthquake led to a revolt against Sparta by its subject peoples, the Helots of Laconia and Messenia. A later Greek historian, Diodorus Siculus, agreed, calling the earthquake “a great and incredible catastrophe,” which was part of a history of earthquakes in the area (Diodorus Siculus, 1946, p. 289). A fault scarp 10 to 12 meters high and about 20 km long passes within only a few kilometers of the site of ancient Sparta. Diodorus is the source of the claim that more than 20,000 Spartan citizens died in the earthquake under collapsing stone walls, as a result of the “tumbling down of the city and the falling of the houses” over a long period. The high number of Spartan casualties is thought by modern historians to be an exaggeration. Nevertheless, the Spartans were in due course obliged to recruit noncitizens as hoplites for their army, and the recruits were less dedicated to the famously disciplined Spartan military code.

Without doubt, the earthquake in Sparta was a factor in the origin of the Peloponnesian wars between Sparta and Athens, which began in 460 bc and lasted until the end of the 5th century. When the Helots, a population subjugated by Sparta, saw that a majority of the Spartans had perished in the earthquake, “they held in contempt the survivors, who were few,” writes Diodorus. The Helots took the opportunity of the earthquake to rebel against the Spartans, who requested help from the Athenians, who in response sent a force of Athenian hoplites to Sparta. However, the Athenian presence in Sparta disturbed the regime. According to Thucydides, the Spartan aristocracy, fearing that the Athenians would make common cause with the Helots, sent them back to Athens, thereby permanently souring Sparta’s relationship with Athens.

Although Sparta emerged victorious from the Peloponnesian wars in 404 bc, it soon began to decline, in comparison to Athens, during the 4th century. Its citizens were greatly outnumbered by noncitizens, as noted by Aristotle, falling from about 8,000 Spartan families in 480 bc to not many more than 1,000 in 371 bc. To what extent this fall-off was due to the 464 bc earthquake, rather than Sparta’s loss of citizen manpower during its long struggle with Persia in 499 to 449 bc, or to fatal flaws in the Spartan social structure, is debatable. However, “there is no room for doubting the causal link between the massive Helot uprising of circa 464, with all that it implied for Spartan foreign and domestic policy, and the immediately preceding earthquake,” writes a leading historian of Sparta, Paul Cartledge. “The doubts concern rather the size of the quake … and its longer-run effects” (Cartledge, 1976, p. 26).

By an odd coincidence, the date of the earliest reported earthquake in ancient Italy is almost the same as that in ancient Greece. In 461 bc, according to the historian Livy, writing in the final decades of the 1st century bc, Rome was shaken by a violent earthquake. Another, in 83 bc, damaged the city’s public buildings and houses, and was treated as an omen of civil war. Throughout its history, up to the present day, Rome and its rulers have suffered from both large and small earthquakes, “and the chronicle of these events is one of the best in the world because of Rome’s long historical record,” according to The Seven Hills of Rome: A Geological Tour of the Eternal City (Heiken et al., 2005, p. 97). For example, two earthquakes in ad 443 and 484 damaged the Colosseum, as is clear from a tablet near the entrance to the building that thanks Decius Marius Venantius Basilius for his generosity in subsidizing its repair. Another major earthquake in 1349 most likely damaged it beyond repair: the southern section of the ancient amphitheatre’s external wall collapsed, while the northern wall remained standing. A seismic study in 1995 of the Colosseum’s foundations, using sound waves to create images of the subsurface structure, revealed that the southern section rests on alluvium—accumulated sediment filling the prehistoric bed of a tributary of the River Tiber that is extinct—whereas the northern section stands on the riverbank, where the ground is older and more stable.

That said, unlike in Sparta, no earthquake in ancient Rome was ever severe enough to shake the stability of the Roman state. Other parts of Italy, such as the Apennine mountains in the spine of the peninsula, and the area around Naples, including Pompeii, suffered from much greater shaking than the capital city—as they do in the early 21st century.

London, United Kingdom, 1750

Britain has a long history of earthquakes, including one in 1580 referred to by William Shakespeare in Romeo and Juliet. Some have caused serious damage and a few fatalities, but none is of significance to the world, except for the earthquakes of 1750, the so-called “Year of Earthquakes”—with five notable ones in England—that caused Londoners to panic and Britain’s natural philosophers to begin the scientific study of seismicity.

The first struck soon after half past midday on February 8, 1750 (according to the Julian, “Old Style,” calendar in use up to 1752). Although small, its epicenter was beneath the capital, apparently around London Bridge. So the city received a considerable jolt. Throughout the City and Westminster people felt their desks lurch, chairs shook, doors slammed, windows rattled, and pewter and crockery clattered on shelves.

To begin with, an earthquake was not accepted as an explanation—so improbable did it appear to be. Instead, there were theories about cannon-fire and exploding powder magazines. Then it was said that Isaac Newton, before his death in 1727, had predicted the jolt by calculating that Jupiter would approach close to Earth in 1750. Within two or three weeks, Londoners began to forget the strange experience.

Almost exactly four weeks after the first shock, at 5:30 am on March 8, came the second. More pronounced, it covered five times the area.

Horace Walpole, man of letters and member of Parliament, was in bed in central London. Three days later, he reported to a friend outside London: “On a sudden I felt my bolster lift up my head; I thought somebody was getting from under my bed, but soon found it was a strong earthquake, that lasted near half a minute, with a violent vibration and great roaring” (Walpole, 1906, pp. 198–199). The president of the Royal Society, antiquarian Martin Folkes, noted that the vibration and noise could not have been that of a passing cart or coach—to which many compared it—because everything was entirely quiet at such an early hour. In west London, near Holland House, a bailiff counting his sheep observed the dry, solid ground move like a quagmire or quicksand, causing much alarm among the animals and some crows nesting in nearby trees.

A slight tremor occurred on March 9, and then came a powerful rumor: a third shock, exactly a month after the second one, would swallow up London. The rumor was started by an army trooper who would eventually be dispatched to Bedlam, London’s lunatic asylum. By April 4, doomsday had somehow advanced to the very next day, and panic took hold. “This frantic terror prevails so much, that within these three days seven hundred and thirty coaches have been counted passing Hyde Park corner, with whole parties removing into the country,” a skeptical Walpole reported from the frontline. “Several women have made earthquake gowns; that is, warm gowns to sit out of doors all tonight” (Walpole, 1906, pp. 202–203).

That Walpole was not exaggerating is confirmed by the “Historical Chronicle” of April 1750 published in the monthly Gentleman’s Magazine. For April 4, it reads:

Incredible numbers of people, being under strong apprehensions that London and Westminster would be visited with another and more fatal earthquake … left their houses, and walked in the fields, or lay in boats all night; many people of fashion in the neighbouring villages sat in their coaches till daybreak; others went to a greater distance, so that the roads were never more thronged, and lodgings were hardly to be procured at Windsor; so far, and even to their wit’s end, had their superstitious fears, or their guilty conscience, driven them.

Part of the blame for the exodus must undoubtedly fall on the activities of religious preachers during March. A leading clergyman, William Whiston, successor to Newton as Lucasian professor of mathematics at Cambridge University, expressed his long-held belief that the end of the world was quite close at hand, as predicted by 99 signals. No. 92 was that there would occur a terrible—but to good men a joyful—earthquake, which would destroy one tenth of an eminent city. But it was the bestselling tract of the bishop of London, Thomas Sherlock, a former vice-chancellor of Cambridge University, which attracted the most attention. He urged his readers to repent, and to ignore “little philosophers, who see a little, and but very little into natural causes … not considering that God who made all things, never put anything out of his own power” (Gentleman’s Magazine, 20, 1750, p. 123).

When London failed to fall, there was a general air of sheepishness in society, followed by a forgetting of the earthquakes, with no attempt to protect London’s buildings from future shocks. Natural philosophers, however, became fascinated by seismicity. By the year’s end, almost 50 earthquake articles and letters—mainly personal narratives by witnesses—had been read before the Royal Society, which were promptly published as an appendix to its Philosophical Transactions.

One of the society’s fellows was John Michell, a Cambridge astronomer who had a remarkable range of interests, including geology. During the 1750s, Michell examined eyewitness reports from England in 1750 and from the terrible Lisbon earthquake in 1755, and analyzed them according to Newtonian mechanics. His important if flawed paper, “Conjectures Concerning the Cause and Observations upon the Phaenomena of Earthquakes,” published in the Philosophical Transactions for 1760, correctly concluded that earthquakes were “waves set up by shifting masses of rock miles below the surface”—although his explanation for this shifting relied wrongly on explosions of steam, as underground water encountered underground fires. There were two types of earthquake wave, he said: a “tremulous” vibration within the earth, followed by an undulation of the earth’s surface—once again he came close to the truth. Despite being a clergyman, and notwithstanding Sherlock’s warning against “little philosophers,” Michell boldly left the divine out of his analysis.

Lisbon, Portugal, 1755

Unlike in England in 1750, earthquake data were collected systematically in Portugal in 1755—for the first time in the history of earthquakes. An official questionnaire was distributed to parishes by the government. For instance: Did you perceive the shock to be greater from one direction than another? Did buildings seem to fall more to one side than the other? Did the sea first rise or fall? How many hands did it rise above the normal? The answers were stored in the national historical archives in Lisbon.

The driving force behind this enquiry was the prime minister, best known under his later title as the first marquess of Pombal, the power behind the religious-minded king José I. The earthquake was the making of Pombal. In its bewildering and frightening aftermath, when the impractical king asked his practical prime minister what he should do, Pombal is supposed to have given the blunt advice: “Bury the dead and feed the living” (Kendrick, 1956, p. 45). As noted by Pombal’s biographer: “He took quick, effective, and ruthless action to stabilize the situation” (Maxwell, 1995, p. 24). This included dumping weighted corpses at sea and burying victims in mass graves under the city, to suppress disease. As a result, the earthquake’s death toll can never be known for certain; it was probably more than 30,000 Lisboetas.

Less immediately, Pombal had to deal with clerical arguments that the devastation was a divine response to the city’s moral failings. The Jesuit clergy had been put in an especially invidious position by the earthquake. If it truly was divine punishment for the sins of Lisboetas, then why had it occurred on All Saints’ Day, destroyed so many religious institutions, and killed so many clergy? Theologically, the earthquake presented a profound paradox.

Voltaire was incensed by such religiosity. A few days after hearing the first news of the Lisbon disaster, he informed a banker friend: “I flatter myself at least that the reverend Fathers, the Inquisitors, will have been crushed like all the others. That ought to teach men not to persecute men, for while some holy scoundrels burn a few fanatics the earth swallows up the lot of them whole” (Paice, 2008, p. 192). But Voltaire also disagreed with the optimistic outlook of secular society, epitomized in a celebrated essay by Gottfried Leibniz that conceived the world we experience to be “the best of all possible worlds.” In Voltaire’s poem about the earthquake, and his later novel Candide, he questioned how either the church authorities or optimist philosophers could possibly justify the destruction of Lisbon. Why not decadent London or Paris? He asked, Why does Lisbon lie in ruins, while in Paris they dance? Jean-Jacques Rousseau, another philosopher, disagreed with Voltaire’s argument against optimism on the reasonable grounds that people in Lisbon had brought disaster upon themselves by choosing to invest in fragile urban buildings and packing them with valuable possessions, which they hesitated to abandon, rather than living in safe, small houses in natural surroundings, from which they could easily escape. And some other thinkers, such as Immanuel Kant, argued that great earthquakes might have hidden benefits. But for the majority of thinking Europeans, including the young Wolfgang von Goethe, Voltaire’s criticism marked the beginning of an irreversible shift away from both religious explanations of natural disasters and the philosophy of optimism.

Pombal certainly had no truck with the Jesuitical account of the earthquake. Soon after it, the Portuguese government issued decrees to forbid priests from stirring up feelings of recrimination and guilt in the population. When the leader of the Jesuits, Gabriel Malagrida, prophesied a second great earthquake in November 1756, Pombal banished him from Portugal. In 1759, the entire Jesuit order was expelled from Portugal and its empire, and any communication between Jesuits and Portuguese subjects was prohibited. In 1760, Portugal and the Vatican broke off diplomatic relations. The year after, Malagrida was subjected to an auto-da-fé in Lisbon by the Inquisition—led by Pombal’s brother—and garroted; his corpse was then burnt at the stake, and his ashes thrown into the River Tagus.

Meanwhile, Pombal saw the destruction of Lisbon as a practical challenge. From the late 1750s, the increasingly dictatorial Pombal pursued the rebuilding of Lisbon, partly funded by imposing an extra 4% import tax (over the strong objections of British merchants in Lisbon). Indeed, “he seized on the disaster as an opportunity for urban development and commercial expansion,” notes Kevin Rozario in The Culture of Calamity. “Employing the absolute power of the monarchy and the dwindling but still fabulous riches of empire, Pombal built a magnificent new metropolis, and he overhauled the economic and political structure of his country” (Rozario, 2007, pp. 17–18). The rebuilt area of Lisbon, known as the Baixa, is still impressive for its grid of wide thoroughfares. Moreover, the new buildings had wood-lattice frames—the so-called Pombaline cage—designed to move with the shaking of the earth: they are one of the earliest examples of earthquake-resistant construction in wood (along with some Japanese pagodas).

The rebuilding process continued well beyond Pombal’s fall from power in 1777 after the death of his royal patron, and the occurrence of two further earthquakes in 1796 and 1801. But then Lisbon suffered again, from a destructive 4-year occupation by the French troops of Napoleon Bonaparte, in which the Portuguese royal family was forced to flee to Brazil in 1808. Although, over the decades, the city had recovered much of its former prosperity, the loss of Brazil as a colony in 1822 was a third serious blow—after the earthquake and the occupation—from which Portugal would never recover. From this time onward, Lisbon was haunted by a sense of deep loss—expressed in the sorrowful Portuguese singing known as fado (“fate”), which began to be heard on Lisbon’s streets from the 1820s. As Charles Dickens noticed on his visit in 1858, for all the color, charm, and exuberance of the city’s street life, the “common people” of Lisbon never seemed to laugh (Dickens, 1858, p. 88).

Caracas, Venezuela, 1812

Brazil’s declaration of independence from Portugal in 1822 after three centuries of colonial rule was part of a continent-wide movement in Latin America toward political independence. In 1800, at the beginning of the Napoleonic wars, virtually all of Central and South America was under Spanish or Portuguese royal rule. By 1825, this dominion had been swept away, except in Cuba and Puerto Rico, leading to the creation of new nations: Bolivia, Colombia, Ecuador, Peru, and Venezuela.

Although the first of the political “earthquakes” in this storm struck Haiti, when it gained its independence from France in 1804, the most influential occurred in the young Republic of Venezuela, whose independence from Spain had been declared in July 1811 under the leadership of General Francisco de Miranda and Simón Bolívar. Just before Easter, on March 26, 1812, a real earthquake severely damaged Venezuela’s capital, Caracas, and some other parts of the country.

By chance, the damage happened to be worst in the areas controlled by Miranda, Bolívar, and the rebels, who called themselves “patriots.” The local Catholic authorities had a field day with this portentous coincidence just before Easter. The archbishop of Caracas thundered that the earthquake was a terrifying but well-deserved punishment for the city’s “patriot” disloyalty to Spain, and invoked the biblical warnings about Sodom and Gomorrah. Bolívar, while organizing republicans to rescue the dying and the dead by digging with their bare hands and removing the victims on makeshift stretchers, stumbled across a red-faced priest berating a frightened crowd. “On your knees, sinners!” the priest shouted. “Now is your hour to atone. The arm of divine justice has descended on you for your insult to his Highest Majesty, that most virtuous of monarchs, King Ferdinand VII!” Legend has it that Bolívar threatened the priest with his sword (Arana, 2013, p. 109).

This religious exploitation of the earthquake in a highly superstitious land was effective at stirring up a rebellion against the republic. In July 1812, Spanish troops retook the country and captured Miranda and Bolívar, who was lucky to be sent into exile in Cartagena. There, Bolívar noted in a key political manifesto: “The earthquake of 26 March was as devastating physically as it was spiritually and can fairly be said to have been the immediate cause of Venezuela’s ruin” (Bolívar, 2003, p. 7). But in the end, Venezuela’s seismic undoing in 1812 would prove to be Bolívar’s making.

In 1813, Bolívar was named commander of an expeditionary force intended to liberate Venezuela. After a sweeping campaign, in six pitched battles he defeated the Spanish, re-entered Caracas, and founded the Second Republic. Although he was soon ousted again, he returned to the fray repeatedly and finally freed Venezuela from Spanish rule after a battle in 1821. The following year, Ecuador was secured after another battle led by a lieutenant of Bolívar. Two years later, two further battles led by him and his lieutenant liberated what then became Peru and Bolivia. Since Argentina had declared independence in 1816, Chile in 1818, and Brazil in 1822—under different leaders—by 1825 almost the entire South American continent was politically independent.

From then until Bolívar’s death in 1830, there were bitter disputes between him and other leaders, including an assassination attempt and his resignation as president of Gran Colombia: the federation of Bolivia, Colombia, Ecuador, Panama, Peru, and Venezuela plus western Guyana and northwestern Brazil. But there can be no dispute about Bolívar’s historic importance as a liberator. Equally clear is the role of the 1812 earthquake in pushing him down this path of liberation, beyond his native Venezuela. As Bolívar recognized, without the earthquake’s devastation of Caracas and other republican cities, and the royalist backlash against the republican cause that followed the quake, the First Republic of Venezuela might have survived, and he would have remained there at its head. In which case, the liberation of northern South America from Spanish colonial rule—lacking Bolívar’s seismic leadership—would probably have taken much longer than it actually did.

Naples, Italy, 1857

Although supernatural explanations of earthquakes were still prevalent during the early 19th century, science had by then made some progress in understanding seismicity, without as yet becoming the discipline of seismology—beginning with the geological paper by John Michell published in 1760. However, Michell’s insights had no influence on the further study of earthquakes until the mid-19th century. Theories about seismic waves were all very well, but first it was necessary to try to measure earthquakes, by carefully examining and classifying the destruction they caused.

The pioneer in this area was Robert Mallet, a brilliant Irish civil engineer and fellow of the Royal Society, who coined the term seismology in 1857. From the 1830s, for over 20 years Mallet collected as much data about historical quakes as possible. His catalogue of world seismicity contained 6,831 listings, giving the date, location, number of shocks, and probable direction and duration of the seismic waves, along with notes on related effects. But his real breakthrough occurred in the field: in Italy, following a devastating shock in an area close to Naples—the province of Basilicata—in mid-December 1857. With as many as 19,000 fatalities, this earthquake was supposed to be the third greatest European earthquake ever recorded, after Lisbon in 1755 and Catania in Sicily in 1693.

Less than two months after the earthquake, with the support of the Royal Society, Mallet was hard at work in Basilicata, travelling on horseback. It was often tough going, with much rain and intense cold at night, necessitating stays in precarious lodgings, apart from a few monasteries and large houses. Even without the devastation of an earthquake, the region was not easy to penetrate. Much of it lay between 1,000 and 1,500 meters above sea level, with higher peaks. Mallet, lacking provincial Italian, had to gather information through interpreters in a variety of dialects, apart from one direct conversation—in Latin!—with a padre.

“At first sight, and even after cursory examination, all appears confusion. Houses seem to have been precipitated to the ground in every direction of azimuth. There seems no governing law, nor any indication of a prevailing direction of overturning force,” Mallet wrote in his official report. Of his general approach he observed:

It is only by first gaining some commanding point, whence a general view over the whole field of ruin can be had, and observing its places of greatest and least destruction, and then by patient examination, compass in hand, of many details of overthrow, house by house and street by street, analysing each detail and comparing, as to the direction of force, that must have produced each particular fall, with those previously observed and compared, that we at length perceive, once [and] for all, that this apparent confusion is but superficial.

(Mallet, 1862, Vol. 1, pp. 35–36)

By assessing each crack of the earthquake’s damage with a trained eye, Mallet compiled iso-seismal maps: that is, maps with contours of equal earthquake damage/intensity (a method still employed, with refinements, to map seismic hazard). Although he placed too much reliance on the direction of fallen objects and the type of cracks in buildings as indicators of earthquake motion, Mallet’s maps allowed him to estimate the center of the shaking and the size of the earthquake relative to other earthquakes. However, he never stumbled on the massive ground ruptures created by the geological fault that slipped in the 1857 quake, which were discovered only in the 1990s. Why did Mallet miss them? One reason may have been that in 1857—decades before great ruptures were observed in Japan in 1891 and California in 1906—no geologist expected to see a surface rupture, because seismic movement was at this time assumed to take place deep underground. Another reason is meteorological: the ruptures may have been hidden by snow.

Nonetheless, Mallet’s report on the 1857 earthquake “constitutes a watershed between descriptive naturalism and the scientific and speculative observation of an earthquake,” according the Notes and Records of the Royal Society (Ferrari & McConnell, 2005, p. 62). Mallet used the new technique of photography, including stereoscopic images, to document the damage in detail. His subsequent lengthy text, with illustrations, was published in a two-volume study, Great Neapolitan Earthquake of 1857: The First Principles of Observational Seismology, in 1862. He also published a catalogue of historical earthquakes that remains a standard reference, and some extraordinary maps of seismic intensity throughout the world’s landmasses that were not bettered until the seafloor measurements of the 1950s. Mallet’s maps gave the first indication that earthquakes cluster in certain belts around the earth. An explanation of why this is so, involving plate tectonics, would have to wait for another century, but in the meantime Mallet’s map directed the attention of geologists to the mysterious global seismic intensity patterns.

San Francisco, United States, 1906

California’s earthquakes were soon a key focus, leading to the identification of the San Andreas Fault in 1895. The world’s most famous earthquake occurred on the fault soon after five o’clock in the morning of April 18, 1906, in two shocks separated by a pause of between 10 and 12 seconds, lasting between 45 and 60 seconds in all. The lower part of San Francisco’s City Hall collapsed almost instantaneously, as did many other buildings. But it was the fire that started after the earthquake had disabled the main water supply that destroyed the city. Despite heroic efforts from firefighters and the dynamiting of buildings to create fire-breaks, the fire burned for three days until it was at last deprived of fuel around seven in the morning on April 21. By then it had devoured 508 blocks and more than 28,000 buildings, extending over 12.2 square kilometers—three-quarters of the city (and about eight times the area destroyed by the Great Fire of London)—and had cost at least $500 million. Combined with the earthquake, the fire left 225,000 people—more than half of the city’s population—homeless.

Politicians and businessmen in the city unquestionably made a sustained effort to blame the devastation on the fire rather than the earthquake, and this belief became firmly established—so much so that in later years people spoke of “before The Fire” and “after The Fire” as historical benchmarks. The denial of seismic truth found a ready response among ordinary San Franciscans for the following reasons. First, building insurance policies generally covered fires, but not earthquakes, which naturally encouraged policyholders to downplay the earthquake. Second, focusing on the fire deflected attention away from any long-term seismic risk to the city, which was good for business confidence. Third, blaming the fire encouraged everyone to rebuild the city as it was before, as quickly as possible, without the need for any expensive changes to the foundations of buildings or any antiseismic structural engineering. San Francisco’s daily newspapers abetted the belief by publishing telegraphic reports of small tremors in the eastern United States, while omitting to report the stronger aftershocks in San Francisco itself.

Indeed, the State Earthquake Investigation Commission received little help from San Franciscans in gathering information on the event. Pressure was exerted on its scientific members by the city authorities and the chamber of commerce to suppress the commission’s findings, as had happened after the city’s previous earthquake in 1868, when a disturbing scientific report was never published. One 1906 commission member, John Casper Branner from Stanford University (which was badly damaged by the earthquake), candidly recalled in 1913: “We were advised and even urged over and over again to gather no such information, and above all not to publish it. ‘Forget it’, ‘the less said, the sooner mended’, and ‘there hasn’t been any earthquake’, were the sentiments we heard on all sides” (Fradkin, 1999, p. 136).

Even the insurance companies connived in the view that the disaster was pyrogenic, rather than seismic, in origin—despite a guarantee that this was against their financial interests. A building that fell in an earthquake, and then caught fire, was not covered by standard fire insurance. Policyholders were well aware of this. After the conflagration, the chairman of the Fire Underwriters’ Adjustment Committee said: “It may be considered an extraordinary thing that in the first 2,000 claims submitted to insurance men in San Francisco, after the earthquake and fire, every man filing a claim swore that his property was uninjured by the earthquake” (Hansen & Condon, 1989, p. 124). In the end, an across-the-board compromise was reached, known as the “horizontal cut,” that is, a percentage discount of the full value of an insurance policy, allowing for the possible effects of earthquake or dynamite damage. Insurance adjusters pushed for the discount to be one third, but after negotiations between the insurance companies and a committee of policyholders, the discount came down to 10%. This became the basis for future statistics that attributed 10% of the overall damage in 1906 to the earthquake and 90% to the fire.

Along with the insurance adjusters came an army of researchers from engineering societies, universities, and government agencies. They stayed for longer, about 3 years, treating downtown San Francisco as a laboratory for studying the structural engineering aspects of the earthquake and fire—like Mallet in the ruins of the great Neapolitan earthquake of 1857. Their aim was to discriminate between damage caused by the earthquake, dynamiting damage, and fire damage, and thereby to provide a basis for rebuilding a more earthquake- and fire-resistant city.

But it was not to be. San Francisco spectacularly rebounded after 1906, but it retained most of its earlier faults. The gleaming new city that emerged from the ashes of the old city was structurally less sound than the pre-1906 city. Within decades, though, its prominence stimulated the birth of the high-tech industrial area, Silicon Valley, southeast of San Francisco and also on the San Andreas Fault. The Californian attitude to earthquakes has always been an ambivalent mixture of embrace and denial.

Tokyo and Yokohama, Japan, 1923

Only in Japan is the protection of urban infrastructure, buildings, and people against earthquakes treated with deadly seriousness. But then, Japan is the only industrialized nation in the world ever to have suffered the almost total destruction of its capital city by an earthquake.

The Great Kanto earthquake of 1923 struck Tokyo and its nearby international port, Yokohama, just before lunchtime on September 1. The cities were subjected to between four and five minutes of shaking, followed shortly after by a tsunami. The energy released was equivalent to some 400 Hiroshima-sized atomic bombs. Charcoal and gas braziers were then cooking the midday meal in a million wooden houses. Soon multiple small fires started in panic-stricken kitchens. Feeding on the congested houses, the fires merged to form terrifying firestorms that burned through the night. By the morning of September 3, 18 square kilometers of Tokyo had been incinerated (one and a half times the burnt area of San Francisco in 1906). At least 140,000 people died.

Little survives of the 1923 earthquake and fire. The evidence was largely destroyed by reconstruction in the 1920s, then obliterated by incendiary bombing of Tokyo during World War II in 1945. Indeed, United States military planners deliberately based their aerial attack on the destruction caused by the post-earthquake fire in 1923.

Also obliterated was a grand plan for reconstruction conceived by the home minister of the imperial government, Goto Shinpei, a former mayor of Tokyo, in the immediate wake of the disaster. He saw the destruction as a blessing in disguise: a chance to clear away Tokyo’s burgeoning slums and remodel the city on a European-style grid as a capital worthy of a great power (like Pombal in Lisbon after 1755). But other Japanese politicians, including the finance minister, were in no doubt that such a grand plan would cost far more than the nation could afford. In practice, two thirds of the reconstruction budget spent between 1923 and 1930 went not on creating a brand-new Tokyo but on improving the roads, canals, and bridges of the old Tokyo, along with a process of “land readjustment.” This entailed a street-by-street negotiation with residents, who had to sacrifice up to 10% of their private land without government compensation in the interests of bettering their city—largely by eliminating narrow alleyways in favor of straight modern roads with pavement. By 1930, Tokyo was officially declared to be reconstructed—even more quickly than happened in San Francisco after the 1906 earthquake.

As for the long-term effects of the Great Kanto earthquake, opinions differ. Although it is certain that the earthquake did not cripple Japan’s economy, it undoubtedly destabilized it. According to Edward Seidensticker in Tokyo Rising: The City Since the Great Earthquake, debts arising from the earthquake played a direct role in a financial panic and bank run in 1927 that led to the resignation of the cabinet and the appointment of a general as prime minister, and he advocated aggressive military interventionism in China. In 1931, Japan invaded and occupied Manchuria.

While it is reasonable to postulate a causal connection between the massive disruption of the earthquake and the eventual declaration of total war by Japan in 1941, it is more difficult to substantiate it. The consensus among historians and political commentators is that the martial law imposed between September 2 and November 15, 1923, gave new authority to the police and the army, which some officers exploited for their own agendas, including the massacre of Korean immigrants. The earthquake “paved the way for political domination by the Japanese military,” writes Richard Samuels. “The military had rallied a vulnerable nation under banners of leadership, social solidarity and change” (Samuels, 2013, p. 55). It was a “turning point,” agrees David Pilling, which “helped crystallise a lurch towards totalitarianism” (Pilling, 2014, pp. 72, 304). The declaration of martial law “gave an already aggressive military new power and stature in Japanese society,” confirms Joshua Hammer. “Many of the officers who rose to positions of authority in the earthquake’s aftermath would play prominent roles in the radical anti-democratic groups that formed in the late 1920s and early 1930s, the same groups that would lead the country to war” (Hammer, 2006, p. 259).

Instead of political influences from the earthquake, Seidensticker—who is best known as a scholar and translator of Japanese literature—discerned subtle, if unproven, cultural influences. In particular, the Japanese passion for cartoon strips and comics with a panel narrative, familiar as manga, dates from the years immediately after the Great Kanto earthquake. “Whether or not their origins can be blamed on the confusion that followed the earthquake, that is where they are,” Seidensticker observed (Seidensticker, 1990, p. 121). Given the undoubted origin of an earlier distinctively Japanese graphic form—namazu-e, the fascinating colored prints that depict an earthquake as a mischievous underground catfish (namazu)—in the events of a great earthquake in the capital city in 1855, a similar kind of artistic and literary phenomenon arising from the 1923 earthquake seems plausible.

Tangshan, China, 1976

However one interprets the Great Kanto earthquake’s impact on Japan, its impact makes a striking contrast with that of a great earthquake in China half a century later. The Tangshan earthquake not only catalyzed the end of a military dictatorship and the rise of a great economic power but also marked the end of China’s violent Cultural Revolution, launched by Mao Zedong in 1966.

In 1976, Tangshan was a sprawling industrial city, 150 km roughly to the east of Beijing. It had a population of over one million people: a thousandth of the whole of China, but responsible for a hundredth of the country’s economic output, it was often said. The city was the home of China’s first modern coal pit, its first standard-gauge railway (built to haul coal), its first steam locomotive, and its first cement works—all of them originally constructed by foreigners in the late 19th century.

But there was to be no foreign help after the earthquake on July 28, 1976. The Chinese government flatly rejected immediate offers of aid from “foreign devils” and drew a cloak of secrecy over the near-total destruction of Tangshan. Nor was any official death toll announced in 1976. It was eventually given as 242,000, but this covered only the city’s residents, excluding migrant workers, visitors, and the surrounding area. The toll for all of these is widely thought to be as high as 750,000.

At this desperate time for the people, the sole high-level Communist party leader to visit Tangshan was Vice Chairman and Premier Hua Guofeng, the designated successor to Mao. Hua made a good impression in Tangshan. His official report, delivered to Mao on August 18, was the last official document Mao read. Mao’s death in Beijing on September 9 immediately intensified an existing power struggle between the government led by Hua, the so-called Gang of Four (led by Mao’s widow), and the followers of Deng Xiaoping, who was still officially in disgrace. Yet without the trauma of the earthquake, both this short-term political struggle, and the long-running national self-destruction of the Cultural Revolution, might well have dragged on for far longer than they actually did, stymieing the much-needed reform of Mao’s failed political and economic policies.

In contrast to Hua, the Gang of Four fatally misread the shock of the earthquake. Not only did none of them visit Tangshan, they also regarded the disaster in rigidly political terms: “Be alert to Deng Xiaoping’s criminal attempt to exploit earthquake phobia to suppress revolution! Solemnly condemn the capitalist roaders who use the fear of an earthquake to sabotage the denunciation of Deng!” (Palmer, 2013, p. 191). Earthquake relief was said to be a diversion from the true revolutionary path, even a cover for counter-revolutionary activities. But in 1976, unlike in 1966 (at the launch of the Cultural Revolution), the nation was disenchanted with revolution, distrustful of radicals, and yearning for economic progress, rather than ideological purity. On October 6, Hua Guofeng, in collaboration with other party leaders, had the Gang of Four unexpectedly arrested and in due course charged with treasonable activities. He also declared a formal end to the Cultural Revolution.

Nevertheless, not until 1979–1980, under Hua’s successor Deng Xiaoping, did rebuilding of houses begin in Tangshan; as late as 1983, two thirds of earthquake survivors remained refugees in their own city. By 1986, reconstruction was essentially complete, in time for Tangshan’s coal mines to fuel factories making cheap toys and plastic goods for an export boom. However, the delay proved advantageous in some ways. Not only were the new buildings—mostly low-rise, without skyscrapers—designed to be earthquake-resistant, they were also designed by professionals, rather than by party officials. Neither Marxist theory nor market capitalism shaped modern Tangshan, in contrast to most other major Chinese cities.

The earthquake itself is remembered in Tangshan in the form of a rather brutalist monument from the 1980s; a 21st-century museum emphasizing both the heroism of individual survivors and the superior wisdom of the Chinese Communist Party; and a memorial wall, finished in 2010, which is inscribed with the names of every known victim. The monument and the museum are unsurprising—but the memorial wall, however familiar this idea may be in the West, is unique in China, in that it recognizes “individual loss, rather than collective sacrifice,” observed James Palmer (Palmer, 2013, p. 247). This contrast perhaps encapsulates the earthquake’s chief historical legacy. In its immediate aftermath, a popular political slogan was: “An Earthquake is an Education in Communism” (Qian Gang, 1989, p. 232). But in the longer term, the Chinese people disagreed with the slogan. They put their faith more in individual effort, expertise, and reward than in collectivization. Deng Xiaoping, though almost as authoritarian as Mao, was astute enough to see that he could not buck this populist trend. He therefore encouraged economic growth, while using it to consolidate the party’s grip on power, which had been seriously undermined during the Cultural Revolution. Thus the Tangshan earthquake, in hindsight, proved to be a birth pang in the painful creation of a wealthy and powerful new China.

Gujarat, India, 2001

An earthquake also played a role in the economic rise of modern India. Although Gujarat, India’s westernmost state, does not lie on a tectonic plate boundary, it has a long history of earthquakes, dating back to the Indus civilization in the third millennium bc. Ironically, one earthquake, in 1819, assisted the British colonization of the region in the 1820s, whereas another, in 2001, enabled the coming to power of a controversial Hindu nationalist who became prime minister of India in 2014.

The January 2001 earthquake struck western Gujarat—the large coastal district of Kutch on the Arabian Sea—close to the border with Pakistan. Geologically speaking, it “qualifies as the most devastating intraplate earthquake in the world,” noted the Bulletin of the Seismological Society of America (Mandal & Rodkin, 2014, p. 2060). Its epicenter lay only 20 km from Kutch’s headquarters: the historic town of Bhuj, which sometimes lends its name to the earthquake. The shaking was felt throughout northwest India, including in the country’s capital, New Delhi, and in much of Pakistan, as well as in western Nepal and even in Bangladesh.

Kutch was then comparatively underpopulated and undeveloped; much of it consists of a vast and empty salt marsh, the Rann of Kutch, bordered in the west by the delta of the Indus River in the Pakistani province of Sindh and in the north by the Thar Desert in India. Even so, the earthquake caused an estimated 20,000 deaths and injured some 167,000 people. Four major towns, including Bhuj, were damaged seriously enough to require the imposition of official emergency measures. Even in the state capital in eastern Gujarat, Ahmedabad—more than 250 km from the earthquake’s epicenter—as many as 50 multistoried buildings collapsed. In total, 783,000 buildings were damaged and 339,000 others destroyed, three quarters of them in Kutch, representing 90% of the district’s housing stock.

Gujarat’s chief minister was rapidly accused of incompetence in his response to the disaster by many of the survivors. In Bhuj, protest demonstrations persistently demanded his resignation. In October, the leadership of his Hindu nationalist party, the BJP, replaced the chief minister with a capable but largely unknown local BJP organizer who had never stood for election outside the party: Narendra Modi.

After a highly contested period for Modi in 2002, when Hindu-Muslim riots broke out and he was forced to resign as chief minister, he returned to power in an election. From 2003 onward, he homed in ruthlessly on the possibilities for using the earthquake’s destruction to develop western Gujarat with a new infrastructure—wide roads and 24-hour electricity—and new industries. Over the next decade, Kutch experienced a spate of industrialization along the lines of the industrial development of China since the 1980s. Modi claimed that the “backward region” demolished by the earthquake had come to resemble Singapore.

This so-called Gujarat model of development was much discussed in the rest of India before the country’s national elections in 2014. More than any other plank of Modi’s election platform, his promise of economic growth led to his victory—despite the suspicion of economists and others that Kutch’s growth owed more to Gujarat’s long history of success in industry and international trade than to the leadership of its current chief minister. In the telling words of an operations manager of one of the oil refineries on the Gulf of Kutch completed in 2006, “Gujarat has always been a major industrial hub, and the attitude of the government has always been to invite industrial activity” (The Financial Times, February 5, 2014).

Modi himself was extraordinarily reticent in his direct public references to the earthquake from 2001 onward. There is no major statement about it from Modi in a detailed study, The Political Biography of an Earthquake (Simpson, 2013), nor any reference to it by Modi in The Modi Effect (Price, 2015), a biography written with its subject’s cooperation. Only after a major earthquake struck Nepal in 2015 did Prime Minister Modi briefly lift his public silence on the Gujarat earthquake. While expressing sympathy for Nepal and offering Indian government relief to India’s stricken neighbor, he simply remarked: “I saw the Bhuj earthquake of 26 January 2001, very closely” (The Hindu, April 27, 2015). Yet, it cannot be disputed that without the happenstance of this earthquake, it is inconceivable that Modi would have stepped into Gujarat’s top political job, and hence into Indian national politics.

Aceh, Indonesia, 2004

In December 2004, an area of seabed off Sumatra of staggering dimensions—similar in size to half of California—slipped about 10 meters. The tsunami it produced created havoc in several countries around the Indian Ocean, causing about 230,000 deaths. The country worst affected was Indonesia, especially the province of Aceh at the northern end of Sumatra, where almost 170,000 people perished. Yet, the tsunami also had a beneficial effect on a decades-old civil war between Aceh and the Indonesian state.

Aceh had a long and independent existence before its absorption as a province of the new nation of Indonesia. In the 13th century, it became a Muslim stronghold, the first in the Indonesian archipelago; and from 1511 it was known as the Sultanate of Aceh. In 1948, the first president of Indonesia, Sukarno, promised Aceh a special provincial autonomy within the new country, with official recognition of Islam. But in reality, after 1950 Aceh was incorporated into the province of North Sumatra as part of the secular and centralized federal republic of Indonesia. A rebellion started in 1953 and lasted until 1959, when Sukarno finally granted separate provincial status to Aceh and wide-ranging freedom in regard to the practice of Islam. Relative peace prevailed until the late 1960s, when Sukarno was forced out of office by an army general, Suharto. Then, in 1971, huge oil and natural deposits were discovered in north Aceh. Exploitation by the Indonesian state in collaboration with an international oil company got underway without the involvement of the Acehnese and without benefit to Aceh. All this re-ignited their resentment and fuelled the birth in 1976 of yet another rebellion, the Gerakan Aceh Merdeka, known as GAM or the Free Aceh Movement, which aimed at complete independence for Aceh.

The Free Aceh Movement fought brutally with the Indonesian army, especially in the 1990s. Hostilities largely ceased in 2000–2003, when the Indonesian government offered Aceh concessions. But when the peace talks failed in 2003, President Megawati Sukarnoputri declared martial law and a state of emergency in Aceh, and sent in large numbers of troops and police. By 2004, not only was the Free Aceh Movement on the defensive, it was losing the support of a large part of the Acehnese, who were desperate for peace. Recognizing this, President Yudhoyono, newly elected in October 2004, offered to reopen secret negotiations with the leaders of the movement. They accepted his offer just four days before the tsunami struck on December 26.

In May 2005, martial law was ended and in August, following five rounds of talks, a peace agreement known as the memorandum of understanding (MoU) was signed. It secured the withdrawal of government troops from Aceh, the disbanding of the military wing of the rebel movement, a greater share for Aceh of revenues from energy, mining, logging, and fishing, greater autonomy for local government and, importantly, the right of the former leaders of the Free Aceh Movement to form local political parties. Most of the MoU’s provisions passed into law in 2006, against resistance from hardliners. Despite many continuing tensions in Aceh—especially over the greater distribution of money to tsunami survivors than to war refugees—peace has prevailed since 2005.

The centrality of the tsunami was specifically noted in the memorandum as follows: “The parties are deeply convinced that only the peaceful settlement of the conflict will enable the rebuilding of Aceh after the tsunami disaster on 26 December 2004 to progress and succeed” (Hyndman, 2011, p. 26). On the MoU’s first anniversary in 2006, President Yudhoyono stated: “The tsunami produced an overwhelming moral, political, economic, social imperative to end the conflict. … I was criticized by those who did not see any benefit from renewed talks with GAM. But I was more concerned about the judgment of history for missing this rare window of opportunity to resolve the conflict” (Hyndman, 2011, p. 26). In summary, writes Jennifer Hyndman in her academic study of the tsunami and its political, economic, and social aftermath: “The tsunami did not cause the peace agreement, but it certainly created conditions that accelerated its signing” (Hyndman, 2011, p. 15).

f*ckushima, Japan, 2011

In March 2011, Japan experienced the largest earthquake in its history, which shook most of the nation for about 6 minutes, causing the skyscrapers of Tokyo to sway back and forth like bamboo in the wind. The epicenter was 24 km beneath the Pacific Ocean. The Pacific plate shifted to the west by an extraordinary 30 to 60 meters. This wrench happened approximately 50 km east of the Oshika peninsula in Miyagi prefecture, on the northeast coast of Honshu known as the Sanriku coast, in the region of Tohoku (which literally means “northeast”). Often known in Japan as the Great East Japan earthquake, the quake is known internationally either as the Tohoku earthquake or as the Tohoku-oki earthquake (to signify that it occurred in the open ocean).

Almost 19,000 people died. A 500-km stretch of coastline was affected, including the total destruction of 129,000 homes, the partial destruction of 255,000 homes, and damage to a further 697,000 homes. About 500,000 people were displaced into temporary shelters, while damage to more than 300 hospitals and other medical institutions forced them to close. The total economic loss was estimated at about 20% of the annual national budget of Japan.

But it was, of course, the destruction wreaked by the tsunami on the f*ckushima Daiichi nuclear power plant that lifted the Tohoku quake into the category of earthquakes that changed the history of the world. The name f*ckushima—which ironically means “Blessed Island” in Japanese—is globally synonymous with nuclear disaster, along with Hiroshima and Nagasaki. As the Japanese prime minister, Naoto Kan, announced on March 13, without exaggeration, “The earthquake, tsunami and the nuclear incident have been the biggest crisis Japan has encountered in the 65 years since the end of World War II” (Guardian, March 14, 2011).

Kan vowed to phase out nuclear power, but his political successors were less convinced. Their vacillation is a microcosm of the overall aftermath of the Tohoku earthquake and tsunami. It has become a rerun of events following the Great Kanto earthquake in the mid-1920s, albeit in a more muted form. To oversimplify a little, three basic political attitudes to the crisis have expressed themselves in Japan since 2011. The first maintains that Japan should “put itself in gear” and move in a new direction, away from its dependence on nuclear power and on the United States. The second takes a more conservative stance, that the disaster was so improbable an event as to be unrepeatable—a proverbial “black swan”—hence Japan should “stay the course,” maintaining business as usual, if more efficiently conducted. The third argues that Japan must turn “back to the future,” by restoring its core values and its essential identity, which were lost sight of in its rush to modernity since the 19th century and to globalization in the late 20th century. Thus far, elements of all three attitudes have made some public headway: for example, in the Kan government’s move to phase out all nuclear power; in the persistence of the central government nexus between business and politics, epitomized by the nuclear power industry; and in the emergence of stronger local government in tsunami-affected Tohoku and some other provinces, as well as the growth of a national volunteer movement from its beginnings in 1995 after an earthquake in the city of Kobe.

There have been significant differences from the 1920s, too. As in 1923, there was increased public admiration for the Japanese military, known as the Self-Defense Forces since their establishment in 1954, as a result of their effectiveness in rescue and reconstruction. (In 2011, for the first time, the military were given unrestricted access to nuclear power plants.) But this approval did not raise serious concerns about the militarization of Japanese society. For it was obvious to everyone that the attitude of Japanese society to war had changed irrevocably between 1923 and 2011. Second, the behavior of the Japanese, both in Tohoku and in the rest of the nation, was exceptionally orderly. Probably because of the round-the-clock coverage of the disaster by television, the Internet, and social media, there was no repeat of the 1923 rumor-mongering that led to vigilantism and the murder of Koreans in Tokyo and Yokohama. Third, there was some growth in environmental awareness, symbolized by the launch of Japan’s first green political party, committed to the abolition of nuclear power.

The long-term effects of the Great Kanto earthquake are still being debated by historians. However, if the history of the 1923 disaster is any guide, it seems that the second of the above attitudes—staying the course—is most likely to predominate in contemporary Japan. As one mainstream Japanese politician, Representative Junya Ogawa, bluntly observed in late 2011: “Only 20,000 people died in Tohoku, but 30,000 Japanese commit suicide each year” (Samuels, 2013, p. 180). The Tohoku earthquake and tsunami had not brought Japan to “a tipping point,” said Ogawa, unlike the trauma experienced by the country at the end of World War II.

Conclusion

Aftershock, a Chinese blockbuster movie made by Feng Xiaogang in 2010, turned the 1976 Tangshan earthquake from an unalloyed natural and human catastrophe—the most deadly natural disaster of the 20th century—into both a catalyst for economic development and a nationally strengthening event: all made possible by the continuing leadership of the Communist Party (although this is implied, rather than propagandized). The film was enormously successful in China, perhaps unsurprisingly given its excellent special effects, human drama, and underlying patriotic message. But the truth of Aftershock’s portrayal of the great earthquake and its aftermath was sufficient to carry considerable conviction even with an international audience.

“Creative destruction”—a phrase first used in 1942 by the economist Joseph Schumpeter in reference to the power of capitalism, but here intended to carry a broader meaning—may be said to be part of the aftermath of almost every great earthquake. Creative destruction takes various forms. It can be political, as in Caracas/Venezuela after 1812, Aceh/Indonesia after 2004 and Tohoku/Japan after 2011. It can be economic, as in San Francisco after 1906, Tangshan after 1976, and Gujarat after 2001. It can be cultural, as in Europe, after the Lisbon earthquake in 1755, which provoked Voltaire and Enlightenment philosophers, and after the Naples earthquake in 1857, which kick-started seismology. One might even argue that the Japanese film director Akira Kurosawa’s formative adolescent experience of the Great Kanto earthquake in Tokyo in 1923—witnessing scenes of lawless mayhem and heaps of human corpses—was vital to the creation of his classic films about the extremes of human behavior and emotion, such as Seven Samurai. “Through it I learnt not only of the extraordinary power of nature, but of extraordinary things that lie in human hearts,” Kurosawa wrote in his autobiography (Kurosawa, 1983, p. 50).

Moreover, creative destruction succeeds to varying degrees. It was perhaps most successful in San Francisco after 1906, when construction, business, and innovation all flourished. It was less successful in Tokyo after 1923, when the opportunity to redesign the city was largely lost, a financial panic ensued, and eventually the army took over the government. And in Lisbon after 1755, it was still less successful; the earthquake led directly to the imposition of a dictatorship, and in due course the economic decline of Portugal and its empire. Yet even in Lisbon, the ruined city was redesigned to a new plan that is much admired to this day. “Disaster spurs reinvestment and creative destruction as long as the source of urban economic strength remains fundamentally unaffected. Capitalism, in this sense, outflanks catastrophe,” noted two urban planners in their conclusion to The Resilient City, a study of how modern cities recover from disaster (Vale & Campanella, 2005, p. 347). The urban economic strength of early 20th-century California was on the rise, whereas that of mid-18th-century Portugal was on the decline.

So, how important are earthquakes in history? The historian Will Durant famously claimed: “Civilization exists by geological consent, subject to change without notice” (according to The Oxford Dictionary of American Quotations). Whereas the scientist and geographer Jared Diamond almost ignored natural disasters, and entirely ignored both earthquakes and volcanic eruptions, in his study Collapse: How Societies Choose to Fail or Succeed (2005). Both Durant and Diamond were too extreme. Earthquakes have played a significant and fascinating part in the politics, economics, and culture of many nations and regions of the world from antiquity to the present day. By studying their history, modern civilization can learn how to coexist more securely and creatively with seismic hazard.

Earthquakes in Political, Economic, and Cultural History (2024)

References

Top Articles
Latest Posts
Article information

Author: Eusebia Nader

Last Updated:

Views: 5904

Rating: 5 / 5 (60 voted)

Reviews: 91% of readers found this page helpful

Author information

Name: Eusebia Nader

Birthday: 1994-11-11

Address: Apt. 721 977 Ebert Meadows, Jereville, GA 73618-6603

Phone: +2316203969400

Job: International Farming Consultant

Hobby: Reading, Photography, Shooting, Singing, Magic, Kayaking, Mushroom hunting

Introduction: My name is Eusebia Nader, I am a encouraging, brainy, lively, nice, famous, healthy, clever person who loves writing and wants to share my knowledge and understanding with you.