Keywords

1 Introduction

In 2015, the historian of science Sverker Sörlin posited the emergence of “cryo-history,” histories of ice and snow, during an epoch of retreating cold. The shrinking of glaciers and reduction of sea ice that accompanied global warming had drawn worldwide attention to the mutability of landscapes sometimes assumed to be frozen and timeless. Sörlin observed that ice had a history not only in a geological sense but in social and cultural terms as well. He called for cryo-histories – derived from kryos or κρυος, Ancient Greek for cold, frost, or ice – that examine human relationships with cold and its manifestations (Sörlin 2015). Sörlin’s essay served as much as an indicator as a harbinger. It revealed how, in response to anthropogenic climate change, scholars in the humanities and social sciences had already been turning their attention to the cold (Piper 2009; Carey 2010; McCannon 2012; Martin-Nielsen 2013; Josephson 2014).

The attention encompassed permafrost, the strata of frozen earth beneath the surface of the planet in high-latitude and high-altitude regions. Textual records of perennially frozen earth date to the sixteenth century; systematic investigation into the phenomenon began during the nineteenth century; and the term permafrost appeared in the twentieth century as an English loan translation of the Russian expression, vechnaia merzlota (вечная мерзлота). The development of knowledge about frozen earth was accompanied by debates about its essence. As scholars within and beyond the natural sciences investigated the phenomenon, they have been drawn variously to its materiality and fluidity. While some have emphasized its behavior and impacts as a fixed object, others have focused on its instability as an idea and as an element of nature. The question of whether and how to engage with the indeterminacy of permafrost therefore runs as an underlying divide in histories of frozen earth.

This divide emerged not simply from the challenges to positivism posed by science studies in the second half of the twentieth century. Whereas positivists held that words, including scientific terms, corresponded to a single observable phenomenon, science studies scholars understood them to contain multiple meanings that depended on one’s interpretive framework (Hess 1997). In the case of permafrost, however, the term’s many referents, even from a positivist perspective, are suggested by its definition: “ground (soil or rock and included ice and organic material) that remains at or below 0 °C for at least two consecutive years” (van Everdingen 1998). Permafrost can refer to soil or rock, with or without ice and organic material; it may be frozen in the sense of hardened and held by ice or nonfrozen in the sense of containing liquid water; and it may have recent or prehistoric origins. Permafrost as an idea therefore overlaps imperfectly with encounters with frozen earth in nature. Moreover, the term and its definition do not capture ways of recognizing and interacting with frozen earth outside of European scientific tradition. The multidimensional quality of frozen earth arises both from the varied substances that constitute it and the distinct contexts in which knowledge about it was produced.

While providing an overview of histories of frozen earth, this chapter argues that historiography must attend to the historicity of permafrost as a scientific idea. This imperative has both practical and philosophical dimensions. Since the term permafrost first appeared in 1943, research into the historical record dating before the mid-twentieth century requires knowledge of other linguistic markers of frozen earth (Kabakchi and Butters 1989). Such markers may appear in not only European but also Uralic, Tungusic, Turkic, Inuit, and Yupik languages. They point to different facets of frozen earth, revealing the phenomenon’s relationships to science as well as to the everyday life and worldviews of diverse cultures.

Histories of frozen earth must recover, interrogate, and perhaps help renew these relationships as part of resisting what climate scientist Mike Hulme has called climate reductionism. As Hulme argued, climate reductionism operates by isolating and elevating climate as the dominant factor determining human survival and flourishing on the planet. But doing so crowds out the unpredictability, dynamism, and contingencies of human agency (Hulme 2011). Moreover, by suggesting a direct, causal relationship between climate change and negative social effects, climate reductionism risks reinforcing what philosopher Kyle Whyte has called crisis epistemology. Crisis epistemology casts dangers such as climate change as unprecedented and urgent in hopes of spurring purposeful, decisive action. By portraying problems and their solutions as novel and certain, however, crisis epistemology justifies measures that may perpetuate historical harms, especially to Indigenous peoples. In contrast, Whyte promotes an epistemology of coordination based on Indigenous values of kinship and reciprocity that considers past legacies and cultivates care, cooperation, and repair (Whyte 2020). Such work, prosaic rather than heroic, is necessary for what feminist scholar Donna Haraway has called “staying with the trouble,” or fostering creative interdependence and survival amid crisis (Haraway 2016). Histories that illuminate the many-sidedness and mutability of frozen earth, including and beyond the twentieth-century idea of permafrost, can contribute to this effort.

This chapter examines histories of frozen earth chronologically, using the development, consolidation, and popularization of permafrost science as a guiding organizational principle. It begins by discussing histories written before the establishment of permafrost science as a discipline. Such histories resembled literature reviews in which naturalists and explorers recounted earlier encounters with phenomena related to frozen earth as they advanced fresh ideas and observations of their own. In this stage, terms for frozen earth had not been standardized, and interest in it was connected to major questions of nineteenth-century geology, especially the theory of ice ages. The chapter then considers histories written during the period of consolidating permafrost science. These often aimed to explain the discipline’s progression and relationships to other fields and to educate the broader public. In this stage, scientists also used history to bolster or contest the reputations of key individuals and institutions and to debate the discipline’s goals and orientation. Finally, the chapter analyzes “cryo-histories” – to recall Sörlin’s term – written amid rising global awareness of anthropogenic climate change and the concomitant thawing of frozen earth. If, previously, histories of frozen earth were written primarily by scientists, in this stage scholars in the humanities and social sciences became increasingly engaged as well. They have explored the meanings of frozen earth in a variety of cultural contexts, Siberian and North American, Indigenous and settler. The historiography sets the stage for broader and deeper exploration of how people have imagined and lived with the dynamic phenomenon of frozen earth on a changing planet.

2 Frozen Earth Before Permafrost Science

Frozen earth coalesced as an object of professional scientific interest over the course of the nineteenth century. In prior centuries, scattered mentions of frozen earth had appeared in the writings of explorers, naturalists, and colonial military governors. During the nineteenth century, European men of science sought to compile and synthesize these nuggets of information in monographs and treatises dedicated to frozen earth. On the one hand, the nineteenth-century texts constitute primary sources about frozen earth research. On the other, they represent one kind of historiography of knowledge about frozen earth. Given their position as both primary and secondary sources, they played an important role in defining and negotiating the parameters of frozen earth as a subject of study. For the first historians of frozen earth, writing history constituted part of scientific investigations. Curiosity about frozen earth was strongly connected to debates and discussions about the possibility of previous ice ages. As they gathered and analyzed the record of observations about frozen earth, they wrestled with the phenomenon’s essence and definition. Their debates lent a range of meanings to the ontology of frozen earth, investing it with indeterminacy. This section does not provide a comprehensive survey of published sources about frozen earth prior to the establishment of permafrost science; rather, it highlights those that delved into the history of knowledge about the phenomenon.

Baltic German explorer Alexander von Middendorff is often credited as being the first scientist to provide an overview and detailed observations of frozen earth. He did so in a series of reports presenting the results of his expedition to northern and eastern Siberia from 1842–1845 (von Baer et al. 1848; von Middendorff 1867). However, Middendorff’s studies were guided and informed by the work of Karl Ernst von Baer. Thus, the historiography of frozen earth arguably begins with Karl Ernst von Baer’s Materials Toward Knowledge of the Imperishable Soil-Ice of Siberia (Materialien zur Kenntniss des unvergänglichen Boden-Eises in Sibirien).

In the history of science, Baer is best known for definitively establishing the existence of eggs in mammals (Sarton 1931). But Baer had interests beyond embryology, including biogeography, meteorology, and geology (Tammiksaar 2000). In the late 1830s, he set out to collect all information about frozen earth available in European sources. These ranged from reports of English expeditions to Arctic North America in the sixteenth century to descriptions of northeastern Siberia by Imperial Russian explorers in the early nineteenth century. With these materials, Baer composed a treatise in which he recounted the history of observations of frozen earth and proposed an agenda for future investigations. He also provided an overview of the phenomenon as he understood it, including its appearances, conditions of existence, geographical distribution, and effects upon rivers and groundwater. Completed by 1843, the text ran to over 200 typewritten pages and represented the first written systematic study of frozen earth. Its bibliography serves as an invaluable guide to sources about frozen earth up to Baer’s time. By historical accident, Baer’s Materials remained unpublished until 2000 in Russian translation and 2001 in the original German (Ber 2000; von Baer 2001). The Russian version includes an appendix with Russian translations of some of the primarily German-language sources that Baer consulted. The German version, edited by historical geographer Erki Tammiksaar, provides a more comprehensive bibliography, but without the extracts.

Baer’s monograph set a precedent for and exerted a hidden influence upon subsequent writings on frozen earth. Firstly, it reveals that connections between science and empire enabled the historiography of frozen earth. As a Baltic German, Baer belonged to the noble class of Estland, Livland, and Courland (present-day Estonia and Latvia), which had been colonized by Germans beginning in the thirteenth century. After the region was annexed by Russia in the eighteenth century, Baltic Germans became part of the Russian Empire’s elite. Called “subcontractors of empire” by one historian, they notably occupied positions in state service, the military, and academia (Sunderland 2014). Baer’s circle in Saint Petersburg, where he took a position as a member of the Academy of Sciences in 1834, included officers of the Imperial Navy of German origin. Navigators like Friedrich von Lütke and Ferdinand von Wrangell straddled the worlds of science and empire, their voyages advancing Russian Imperial claims while contributing to natural history and cartography (Daum 2019). The phenomenon of perennially frozen earth came to Baer’s attention through Wrangell, who encountered it while surveying the northeastern coast of Siberia in the early 1820s. Subsequently, Baer organized an expedition to eastern Siberia – ultimately carried out by another Baltic German, Alexander von Middendorff – to investigate frozen earth in situ. The initiative was approved by the Imperial government due partly to the relevance of frozen earth to mining and agricultural colonization (Sukhova and Tammiksaar 2015). Baer relied on intelligence from previous expeditions to write his monograph, which had the goal of informing and guiding Middendorff’s investigations. Activities tied to colonial expansion furnished inspiration and sources for Baer as well as those who built on his work.

Baer’s monograph also posed key scientific puzzles related to frozen earth that were subsequently elaborated upon and debated by scholars of both earth and human history. Baltic Germans’ transnational connections informed these intellectual concerns. Like others of his estate, Baer studied at German-speaking universities and published and corresponded in German, one of the three main scientific languages of nineteenth-century Europe (Whelan 1999; Gordin 2015). He studied at Würzburg and lived for 17 years in Prussia, where he became acquainted with the likes of Alexander von Humboldt and Louis Agassiz (Tammiksaar 2000). Plugged into an international network of scholars, Baer situated frozen earth in the context of new theories that were transforming understandings of geohistory. These included, foremost, the theory of ice ages advanced by Agassiz, which Baer connected to the enigma of extinct mammoths as well as frozen earth. Since at least the ninth century, mammoth tusks had been found and traded throughout Eurasia because of their ivory, a luxury commodity. By Baer’s time, Georges Cuvier had posited that mammoths, parts of whose skeletons and even carcasses could be encountered in Siberia, were “a lost elephant species” – extinct (Cohen 2002; McKay 2017). Central to these debates was a nearly complete mammoth skeleton retrieved from Arctic Siberia by Imperial Russian naturalist Michael Friedrich Adams that had been found with skin and hair (Adams 1808; Woods 2023). What caused the mammoths’ extinction, and how did their parts remain so well preserved? Given that elephants were known to inhabit hot climates in Africa and Asia, how did mammoths end up in cold Siberia?

The nature of frozen earth and the process of its formation constituted part of the puzzle of how mammoths were preserved and possibly transported. At issue was the precise situation at the mouth of the Lena River, along the Arctic coast of Siberia, where the mammoth retrieved by Adams had been located. Adams, relaying the testimony of Osip Shumakhov, the Tungus hunter who discovered the mammoth, suggested that the mammoth had been encased in ice (Tilesio 1815). Such a situation supported a prevailing theory, the drift hypothesis, that hefty objects such as boulders could be transported over long distances, carried in ice floes propelled by water (Krüger 2013). Perhaps mammoths had also been transported from warmer climes via such a mechanism. Drift hypothesis derived from diluvialism, a counterpoint to ice age theory that posited that the earth’s features had been shaped by powerful floods.

On the other hand, subsequent reports by Arctic explorers suggested that mammoths had perished in place rather than being carried from afar. Members of an Imperial Russian expedition in 1815–1818, led by Baltic German navigator Otto von Kotzebue, found an abundance of mammoth tusks on the coast of Alaska. The tusks appeared to have fallen out of cliffs containing massive deposits of ice beneath a layer of soil. Kotzebue and Adelbert von Chamisso, a naturalist on the expedition, referred to the formations variously as ice-ground (Eisgrund), ice-rocks (Eisfelsen), and ice-mountains (Gebirgsart [Eis], Eisbergen) (von Kotzebue 1821; von Chamisso 1821). To Baer, such masses of ice potentially served as evidence of the extensive glaciation posited by Agassiz’s ice age theory. However, later explorers of the area – called Eschscholtz Bay after Johann Friedrich von Eschscholtz, another naturalist accompanying Kotzebue – contested the impression of buried glaciers of pure ice. After visiting Eschscholtz Bay in 1826, British explorer Frederick William Beechey observed cliffs consisting predominantly not of ice but of frozen mud and gravel (Beechey 1831). Eager for more detailed and reliable descriptions, Baer organized expeditions to Siberian sites of mammoth remains by not only Middendorff but also geologist Friedrich Karl Schmidt (von Baer 1866). They attested that the mammoths had been buried in frozen clay and sand rather than encased in pure ice. Their observations supported the view that the creatures once inhabited the far north, their carcasses having been preserved by the frozen earth (Schmidt 1869, 1872). The ice-rich frozen earth observed by the Kotzebue expedition, the mechanism of its formation, and its relationship to mammoths and the ice ages remained a source of scientific debate (Shur et al. 2022). By linking these together, Baer articulated a puzzle that persisted in the history and historiography of science.

Finally, Baer’s Materials carries historiographical weight because it introduced terminology for frozen earth that was subsequently debated, appropriated, and obfuscated. Baer advanced two terms related to frozen earth: Boden-Eis and Eis-Boden. Boden-Eis referred to deposits of ice in the earth, ranging from soil or rock congealed by ice to the ice cliffs observed by the Kotzebue expedition. By Eis-Boden, however, Baer meant that portion of the earth’s crust where the balance of heat exchange between air and land sustained the ongoing presence of Boden-Eis. Whereas Boden-Eis referred to the substance of frozen earth, which required the presence of ice, Eis-Boden referred to a space, anticipating the twentieth-century concept of cryosphere discussed below.

Because Baer’s Materials remained unpublished, his distinction between Boden-Eis and Eis-Boden as well as his spatial understanding of Eis-Boden became forgotten. Middendorff briefly cited and excerpted Baer’s Materials (von Middendorff 1875), and those excerpts were cited by another researcher of frozen earth, Baltic German explorer Eduard von Toll (von Toll 1895). Both writers used Baer’s terms while altering their meanings. Middendorff used Boden-Eis to refer only to large, visible deposits of ice in the earth, which Toll proposed to call Steineis in lieu of Boden-Eis. As for Eisboden, Middendorff established its meaning as soil and rock with a negative temperature, even if ice was not present or visible, which Toll and subsequent writers followed. Instead of denoting a space, Eisboden became the substance of frozen earth, which subordinated the materiality of ice and the notion of a space of heat exchange. This shift led to a fixed rather than dynamic understanding of the phenomenon that became controversial and consequential in the next phases of the historiography of frozen earth.

3 Disciplinary Histories

While nineteenth-century histories of frozen earth revolved around the puzzle of ice age theory, during the twentieth century, economic development provided the dominant context for writings about frozen earth. As states pursued mineral extraction, agricultural settlement, and industrial construction in the Eurasian and North American Arctic and sub-Arctic, frozen earth became perceived as an obstacle. The strategic importance of these projects to countries such as the Soviet Union and the USA prompted governments to nurture the institutionalization of a science of frozen earth. Organizations such as the USSR Academy of Sciences and the US Army Corps of Engineers created institutes and laboratories dedicated to the phenomenon (Obruchev 1947; Wright 1986). Simultaneously, scientific terminology for frozen earth became consolidated through publications and national and international conferences. The English term permafrost was coined from the Russian vechnaia merzlota in 1943, and the first International Conference on Permafrost took place in 1963 (French and Nelson 2008; Dietz 1963). All of these political and intellectual developments shaped scientific understandings of frozen earth into the twenty-first century.

The combination of state interest and the professionalization of science invested histories of frozen earth with political dimensions. These histories were written primarily by practitioners of the new discipline of permafrost science as they strived to define and celebrate their field, establish intellectual genealogies, and sustain funding. Partly what made their narratives political was their explicit or implicit promotion of the strategic economic and military imperatives that motivated state support of permafrost science. Because permafrost science both served and benefited from development projects, scientists and engineers did not interrogate their colonial implications, framing them rather as impetus for scientific progress (Sumgin and Demchinskii 1938; Legget 1954; Brown 1970). Some histories, however, intervened additionally in politics within the discipline of permafrost science itself (Meister and Saltykov 1958). As permafrost science was institutionalized, individuals set forth competing terminologies and frameworks for the discipline while advancing their careers. Their professional, intellectual, and sometimes personal disagreements were exacerbated by different understandings of the essence of frozen earth. Frozen earth’s indeterminacy therefore persisted from the nineteenth to the twentieth centuries. Previously, as frozen earth emerged as a scientific object, its indeterminacy was revealed by the multiple names given to the phenomenon. Now, as a field of science was established around frozen earth, its indeterminacy was revealed by its contested definitions.

As the society that pioneered the science of frozen earth, the Soviet Union also witnessed the most politically charged debates about the phenomenon’s essence. These debates centered on what to adopt as the most accurate scientific term for and definition of frozen earth, revealing disagreement about its ontology. Disputes about terms and definitions also carried broader implications for the goals and methods of the emerging discipline and its position in relation to other earth sciences. Should frozen earth be understood as a concrete body and aggregate structure whose behavior could be tested, measured, and modified for the purposes of stable engineering (Sumgin 1934)? Or should it be understood as a continually evolving condition or process responsive to changes in its surrounding environment (Tolstikhin 1935)? Furthermore, should frozen earth be defined and classified according to its temperature, material composition, age, or spatial location (Parkhomenko 1938; Sumgin 1938)?

Soviet scientists first debated these questions as they and the country as a whole were subjected to enormous pressure to transform an agricultural society into an industrial power. Industrialization and agricultural collectivization formed part of the program of building socialism under the dictatorship of Josef Stalin during the 1930s. They unfolded amid tensions with fascist regimes, human-made famines, and campaigns of repression in which people accused of betraying the party-state were arrested, sentenced to forced labor, and executed. Such a political environment distorted and raised the stakes of scientific debate. On the one hand, the political culture of the ruling Communist Party encouraged scientists to engage in critique and discussion (Kojevnikov 1998; Pollock 2006). On the other hand, it fostered intolerance of pluralism and eagerness to identify scapegoats for failures. Consequently, scientific discussions about frozen earth became enmeshed in accusations of sabotage and individuals’ alleged lack of patriotism (Elenevskii 1937; Obruchev 1940).

The focal point of Soviet scientists’ debate about terminology was the concept of vechnaia merzlota, the origin of the English term permafrost. Vechnaia merzlota was promulgated by Mikhail Sumgin, who published an influential monograph in 1927 that contained the expression in its title (Sumgin 1927). Sumgin spearheaded the USSR Academy of Sciences’ research program on frozen earth, organizing conferences, editing a journal, and becoming recognized as the founder of the science of frozen earth (Obruchev and Chekotillo 1943). His promotion of vechnaia merzlota enabled the term to gain traction in scientific and popular discourse (Sumgin 1935).

But Sumgin did not invent the expression. Rather, vechnaia merzlota emerged from a longer history of practical engagement with frozen earth in the Russian Empire. The word merzlota first appeared in print in an 1838 issue of Mining Journal in connection with strata of frozen earth encountered while searching for gold deposits in Siberia (“Nabliudeniia” 1838). Likewise, vechnaia merzlota first appeared in a report published in 1873 by Russian geographer Petr Kropotkin about an expedition to survey a route to Siberian gold mines. While fulfilling an economically significant task, Kropotkin made geological observations of the territory that brought him into contact with frozen earth (Kropotkin and Poliakov 1873). Then, starting in the 1890s, the Russian Imperial government undertook construction of the Trans-Siberian railroad with a broader vision of bringing agricultural development to Siberia (Marks 1991). Vechnaia merzlota appeared increasingly in works of engineering (Mushketov 1895; Boev 1897; Matseevich 1907; Passek 1911; Bogdanov 1912; L’vov 1916) and soil science (Polynov 1910; Prasolov 1911; Sukachev 1911) as people sought to understand frozen earth for the sake of economic projects. Therefore, alongside research into ice age theory written primarily in German language, the Russian Empire also produced a literature on frozen earth in Russian language oriented to economic development. Vechnaia merzlota as a concept emerged from the latter trend, which persisted as the Russian Empire was replaced by the Soviet Union. By promulgating vechnaia merzlota, Sumgin served as a bridge across the revolutionary divide.

The problem lay in the fact that other scientists found both the term vechnaia merzlota and Sumgin’s definition of it inaccurate and misleading. They argued that the adjective vechnaia – “eternal” – did not correspond to the fluidity, transience, and evolution of natural phenomena, especially on geological timescales (Baranov 1933; Tolstikhin 1935). They also objected to the ambiguity of the noun merzlota, whose multiple meanings rendered it imprecise. Depending on usage, critics pointed out, merzlota could refer to frozen soil, a substance; the condition of frozenness; or a process of freezing that encompassed the crystallization of water (Meister and Shvetsov 1955). Beyond the expression itself, Sumgin provoked controversy with his definition: soil or ground that maintained a negative temperature continuously for at least 2 years and up to tens of thousands of years (Sumgin 1937). Thanks to Sumgin’s initiatives to establish a science of frozen earth, his definition was not only adopted in the USSR but also spread to the USA. It still forms the basis for the definition of permafrost today.

Despite its spread and adoption, however, Sumgin’s definition was continually contested. From the 1930s through the 1950s, different Soviet scientists, including Ivan Baranov, Nestor Tolstikhin, Sergei Parkhomenko, and Petr Shvetsov, raised the same core objections. Firstly, Sumgin defined the substance of frozen earth as soil (pochva) or ground (grunt). But his critics argued that soil and ground belonged to different categories of concepts. Soil referred to specific earth materials, whereas ground referred generally to that part of the earth’s surface that served as a foundation for construction, independent of its material content. Moreover, soil extended a few meters from the earth’s surface, whereas frozen earth was known to extend to depths of over 100 m. Any definition of frozen earth must therefore include rock.

Secondly, Sugmin defined frozen earth according to its negative temperature. But his critics argued that whether or not something was frozen depended not on negative temperature alone. Rather, it depended on the presence of ice, the phase change of water from liquid to solid, which crucially changed the characteristics of the substance in question. In nature, the presence of minerals and impurities in soil and rock meant that water sometimes remained liquid even at some negative temperatures. Thus, even if soil or rock had a negative temperature, it could not be said to be frozen without the presence of ice.

Finally, Sumgin distinguished frozen earth according to the period of its existence, with 2 years as the minimum for characterizing “eternal” frozen earth. But his opponents argued that the length of time for which soil or rock remained frozen depended on other factors. These included its spatial location and material composition as well as its surrounding environment and climate, all of which affected the degree to which it absorbed and radiated heat. Definitions and classifications of frozen earth ought to be based on such primary, independent variables rather than a secondary, dependent variable such as its period of existence. Sumgin’s critics also pointed to the incongruity of labeling frozen earth as “eternal” if it had existed for only 2 years (Baranov 1933; Tolstikhin 1935, 1941; Parkhomenko 1938; Shvetsov 1955).

In lieu of the misleading vechnaia merzlota, Sumgin’s critics developed an alternative view of frozen earth that highlighted its dynamism and connection to the rest of the planet. They took inspiration from the concept of the cryosphere advanced by Polish scientist Antoni Dobrowolski in his 1923 work The Natural History of Ice (Dobrowolski 1923; Zukriegel 1935). Dobrowolski drew attention to ice as an important yet underappreciated substance on the planet, ubiquitous and uniquely influential in both shaping the physical environment and conditioning earthly life. Ice could be found on land, as with glaciers; on water, as with sea ice; and in the air, as in the form of tiny particles. Compared to other solid substances, ice had a low melting point, well within the range of temperatures near the earth’s surface. This property made it highly unstable, perpetually melting amid warmer temperatures and growing amid cooler temperatures. The instability of ice in turn affected the surrounding world, sculpting the terrain and creating or blocking passages across space on the macro- and micro-scales. Some changes had a seasonal rhythm while others took place across longer periods, depending on the latitude and altitude of ice. The spatial distribution of ice varied according to its location along the earth’s axis: nearer to the poles it was found at and below the surface, and toward the equator it was found at high elevations. This space where ice existed and exerted influence constituted an envelope around the planet called the cryosphere.

According to this framework, frozen earth must be seen as not an isolated phenomenon but part of the cryosphere, the realm of ice. Frozen earth comprised the intersection of the cryosphere and the lithosphere, the earth’s crust or outer shell; it might therefore be called the “cryolithozone.” Soviet scientists tied the concepts of cryosphere and cryolithozone to a tradition of geographical thought originating in the Russian Empire. They traced the idea of an envelope of the earth characterized by ice and snow from Mikhail Lomonosov in the eighteenth century to Vladimir Vernadskii in the twentieth. Dobrowolski, a Polish scientist born in the Russian Empire, could be seen as belonging to this intellectual genealogy (Shvetsov and Dostovalov 1959).

Advocates of the cryosphere framework distinguished their physical-geographical view of frozen earth from the engineering and soil science views that gave rise to Sumgin’s vechnaia merzlota. But although its origins differed from the geotechnical view, the cryosphere framework was also shaped by the needs of empire, specifically continental empires. As historian of science Deborah Coen has argued, an intellectual orientation toward thinking across scales emerged as a characteristic of earth science in nineteenth-century Eurasian continental empires (Coen 2018). Such an intellectual orientation was connected to two imperial needs: the need to identify and develop natural resources, and the need to tie diverse peoples and places into one political whole. Both goals required recognizing local diversity – whether in the social realm of language and religion or the physical realm of minerals and climate – within a broader unity. The cryosphere framework, which Soviet scientists traced to patterns of thought in the continental Russian Empire, similarly viewed diverse local manifestations of ice as part of a planetary envelope. Moreover, proponents of the cryosphere framework took interest in the technical challenges of operating on and working with ice and compounds of ice and earth. They also sought the unification and standardization of knowledge in central institutions, whether on the national or international level.

An analogous debate between advocates of vechnaia merzlota and advocates of a cryosphere framework played out in North America. Sumgin’s concept of vechnaia merzlota was promulgated by Siemon Muller, a Russian-American geologist who wrote a foundational English-language monograph on frozen earth based on Russian-language literature. The monograph was commissioned during World War II by the US Army, which encountered frozen earth as an obstacle while building military infrastructure in Alaska and Canada. Muller drew extensively on the work of Sumgin and his colleagues in the USSR Academy of Sciences. He adopted Sumgin’s idea of vechnaia merzlota, including its definition, while giving the Russian expression an English translation, permafrost, shorthand for “permanently frozen ground” (US Geological Survey 1943; Muller 1947).

Although the term was quickly accepted among military engineers, it provoked objections among scientists, notably geologist Kirk Bryan (Barnes 1946; Bryan 1946b). Bryan, like Sumgin’s critics, noted the inaccuracy of suggesting permanence as a characteristic of frozen earth. Recognizing the dynamism of frozen earth, he advocated for terminology that more effectively accommodated the processes of its formation and evolution. To this end, frozen earth ought to be called pergelisol, from Latin roots per (“throughout”), gelare (“to freeze”), and solum (“soil”). From this name, the term pergelation, describing the process of its formation, could be derived. Situating frozen earth within the broader effects of the freezing and thawing of water and ice on the earth’s surface, Bryan proposed to call its study “cryopedology” (Bryan 1946a). His view echoed the cryosphere framework and the idea of “geocryology” as the study of frozen earth developed in the USSR by Sumgin’s critics (Meister and Saltykov 1958).

Despite objections to both, the terms vechnaia merzlota and permafrost stuck. Various circumstances prevented a fuller, transnational discussion of alternative scientific names for frozen earth. These included political obstacles to scientific exchange during the early Cold War, language barriers, and the primacy of geotechnical concerns over theoretical questions. Beginning in the 1960s, however, interest in the cryosphere framework was revitalized by growing research into climate science. Glaciologists became increasingly aware of the importance of ice in regulating the earth’s climate system. They saw the need to understand its behavior in relation to not only the hydrosphere but also the atmosphere and lithosphere. Appreciating the planetary significance of ice drew attention to value of the cryosphere as an organizing concept (Jones 2008; Barry et al. 2011). The cryosphere framework received validation in 2008, when the International Association of Cryospheric Sciences was established within the International Union of Geodesy and Geophysics. In the twenty-first century, some scientists continued to aspire to integrate frozen earth into the cryosphere framework (Dobiński 2006, 2011).

4 Cryo-Histories

Histories of frozen earth entered a new phase in the twenty-first century. From investigating the ice ages and boosting Arctic industrialization, writings about frozen earth became concerned primarily with understanding anthropogenic climate change. In 2000, struck by humanity’s impacts on the planet, scientists Paul Crutzen and Eugene Stoermer posited a distinct geological epoch defined by human activities that they called the Anthropocene. The Anthropocene was characterized by a massive expansion in humans’ use of energy through the burning of fossil fuels (Crutzen and Stoermer 2013; Steffen et al. 2007, 2011). By elevating the concentration of greenhouse gases in the atmosphere, humans were causing average temperatures to rise, altering the Earth’s climate system. As a result, rising sea levels, retreating glaciers, and reduced sea ice as well as heat waves, drought, and tropical storms were bringing harm to human and beyond-human life (Pörtner et al. 2022). The concept of the Anthropocene and the realities of global warming prompted scholars beyond the sciences to rethink humanity’s relationship with the rest of nature (Chakrabarty 2009).

This rethinking included frozen earth, and it drew in humanists and social scientists, not only permafrost scientists themselves. As historians, anthropologists, human geographers, and political scientists investigated frozen earth, they foregrounded its human dimensions. Their engagement with frozen earth contributed to the emerging fields of Arctic social sciences and environmental humanities. On one hand, they explored the intellectual history of frozen earth, the ways that people have known, imagined, and made sense of the phenomenon. On the other hand, they explored cultural practices surrounding frozen earth, the ways that people have interacted and built livelihoods in connection with frozen earth. Rather than take for granted the progress of scientific knowledge or state-sponsored economic development, humanists and social scientists critically interrogated such processes. They were especially attuned to the colonial aspects of the history of science and modernization. When it came to frozen earth itself, they emphasized its pluralism and dynamism. They uncovered multiple essences of frozen earth that stemmed from diverse understandings of the phenomenon. They also emphasized its fluidity rather than its eternity or permanence. Such an emphasis coincided with rising global awareness of the thawing of frozen earth due to climate change (Schiermeier 2001; Blake 2010; Gillis 2011). In highlighting the variability and heterogeneity of frozen earth, humanists and social scientists extended the framework of cryology that Dobrowolski advocated and that Sumgin’s critics embraced. Their cryo-histories approached frozen earth as situated and three dimensional, focusing on local and planetary perspectives while critiquing the territorial gaze of the industrial project of conquering nature.

New histories of frozen earth foregrounded the political and social context in which knowledge about the phenomenon developed. They highlighted the historical processes of resource extraction, infrastructure construction, and military confrontation that motivated people’s encounters with and investigations of frozen earth. In North America, these included the Klondike, Nome, and Fairbanks Gold Rushes and the building of the Canadian-Alaskan Military Highway during World War II (Nelson 2011; Cysewski 2013). They also included the creation of the town of Inuvik for supporting oil and gas development in the Mackenzie Delta, which transformed the lands and livelihoods of Inuit peoples (Stuhl 2016). In Eurasia, scholars emphasized the importance of the Gulag, the Soviet forced labor system, as a powerful sponsor of science and engineering (Khlevnyuk 2003; Siddiqi 2015). Administered by the Soviet interior ministry, which included the secret police, the Gulag functioned as an instrument of not only political repression but also economic development. Although the Gulag operated throughout the USSR, its largest enterprises were located in the country’s far northern and eastern regions, including ones underlain by frozen earth. In these lands, prisoners built railroads, mined coal and gold, and felled trees. To support these projects, knowledge about engineering in frozen earth environments was needed. A key site for researching frozen earth was Vorkuta, the center of a labor camp for mining coal in the Pechora river basin in the far north (Markova and Rodnyi 1998; Barenberg 2014). The science of frozen earth, scholars demonstrated, was intertwined with colonialism and state violence.

While exploring the external, social factors that have shaped frozen earth research, scholars shined fresh light on the phenomenon itself. A history of concepts approach revealed that permafrost constituted only one of multiple ways of understanding frozen earth and recovered other terms, including ones discussed in this chapter (Chu 2020). The different names given to frozen earth and changing motivations for studying it revealed the phenomenon as having many dimensions rather than a single essence. They suggested the need to pluralize rather than flatten knowledge about frozen earth and to incorporate local interactions and perspectives. As Arctic social scientists investigated local practices, they emphasized the relational, rather than fixed, materiality of frozen earth. The materiality of frozen earth depended on myriad local environmental variations, including human behaviors. Human activities and choices, including travel, settlement, construction, and design, altered the thermal conditions that influenced the formation and disappearance of frozen earth (Cho 2021). The materiality of frozen earth also depended on human perceptions of time and the rhythms of human lives. Taking temporality into consideration, frozen earth constituted a solid fluid, solid over shorter time scales but fluid over longer ones (Krause 2022). Social science scholarship therefore echoed the cryosphere framework’s emphasis on the dynamism of frozen earth. But rather than focusing solely on physical substances and systems, it integrated human societies into analyses of nature.

Anthropologists have argued that appreciating the inherent dynamism of frozen earth requires recognizing alternative, non-Eurocentric ways of knowing. They have explored how communities that have lived for centuries in regions with frozen earth built livelihoods connected to the inherent variability of the phenomenon. Such communities included the Sakha and the Eveny of northeastern Siberia. Sakha historically herded cattle and cultivated hay, moving between summer pastures and winter homes. Eveny domesticated reindeer, historically as transport for nomadic hunting and more recently, since Soviet times, for producing meat (Vitebsky 2005). Both communities long understood the fluidity of frozen earth, depending on it for their sustenance and developing worldviews to adapt to it.

Rather than permafrost, for example, Sakha spoke of alaas, which recognized the impermanence of frozen earth. Alaas formed when ice-rich frozen earth thawed, creating a depression that filled with water to become a lake. The lake then dried or drained, and in its wake appeared grassland that Sakha used as pasture for their herds and meadow for haymaking. Sakha not only observed patterns in the changing landscape, but also actively encouraged them: to encourage the formation of alaas, especially its grassland, they selectively burned vegetation to thaw the underlying frozen earth and dug canals to drain the lake. They understood alaas to be a source of life and cultivated a spiritual relationship with the land that encompassed rituals of gratitude and reciprocity (Takakura 2010; Crate 2021).

Eveny also viewed changeability as a characteristic feature of the frozen earth landscape with both benefits and disadvantages. When frozen earth thawed, water from its melted ice replenished the rivers and streams that mobile Eveny used to navigate the land. Yet frozen earth could also block and divert the passage of water, leading to it unexpectedly erupting, flooding, and freezing over large areas and obstructing movement. Eveny observed that the lakes that formed from the thawing of ice-rich frozen earth shifted with strong winds, creating an uneven landscape of shallow depressions. The depressions resulted from the fact that, as pools of water moved, the energy contained in them promoted further thawing of underlying frozen earth. Although moving lakes contributed to further thawing, their movement enabled frozen earth to re-form in the places they vacated. The transitory quality of frozen earth supported the animist worldview of the Eveny in which nonhuman nature is also capable of sentience. For Eveny, the land constituted part of a web of interaction and dependence, continually responding to changes and triggering further changes in the environment. Survival meant accepting and adapting to change rather than expecting permanence (Ulturgasheva 2022).

Attention to the inherent variability of frozen earth has contributed to critiques of the concept of the Anthropocene. As geologists debated the stratigraphic requirements for officially recognizing the Anthropocene as a new epoch in earth history, others have debated the concept’s political implications. For scholars in the humanities and social sciences, the Anthropocene concept valuably highlighted the enormity of human impacts on the planet, including but not limited to global warming. But it also obscured the political and social processes – especially capitalism, colonialism, and racism – that structured the choices and behaviors that led to those impacts (Haraway 2015; Barry and Maslin 2016; Mirzoeff 2018; Yusoff 2018). The Anthropocene concept, scholars argued, collapsed the human species into a universal mankind, homogenizing the unequal harms that have been done as well as the unequal responsibilities. By doing so, it encouraged reliance upon the same processes to solve the problems they created. From the point of view of this critical perspective, frozen earth, too, had been absorbed into a universalist, homogenizing pattern of thinking. Such thinking portrayed the thawing of frozen earth as novel, irreversible, comprehensive, and catastrophic, requiring salvation in the form of bold technological interventions. In actuality, the formation and disappearance of frozen earth were characterized by heterogeneity and discontinuity (Wrigley 2023). The discontinuities of frozen earth created dangers but also opened possibilities for survival rather than salvation.

5 Conclusion

To conclude this tour through the historiography of frozen earth, we might highlight three key claims. Firstly, human engagements with frozen earth invested the phenomenon with multiple essences and multiple names, of which permafrost was but one. Nineteenth-century terms like Boden-Eis, Eisboden, Eisfelsen, and Steineis variously highlighted the substances of ice, soil, and rock that constituted frozen earth. Twentieth-century terms like “cryophilic rocks” and “cryolithozone” drew on the framework of the cryosphere advanced by Dobrowolski. Such a framework recognized dynamism as a characteristic feature of ice resulting from its constant exchanges of heat and energy with its surroundings. This embrace of the changeability of ice differed from the emphasis on stability suggested by permafrost and its definition. But specific historical circumstances enabled the idea of permafrost as ground that maintains a negative temperature for at least 2 years to gain acceptance. The historiography of frozen earth reveals both how permafrost became the dominant way of understanding frozen earth and how alternatives emerged, faded, and persisted.

Secondly, the historiography of frozen earth was shaped by colonialism and its legacies. Interest in the phenomenon was nurtured by connections between science and empire, including both external and internal colonization. States like the Russian Empire sponsored expeditions to Arctic and sub-Arctic regions in hopes of discovering and exploiting natural resources. Such expeditions generated incidental observations and descriptions of frozen earth. As exploration segued into colonization, industrialization, and militarization of peripheries, states sponsored systematic research into frozen earth. Empire- and state-building ambitions influenced ideas about and approaches to studying frozen earth, but their influence, like colonialism itself, had different facets. On one hand, they encouraged a utilitarian approach to frozen earth for the sake of ensuring stable engineering. This approach consolidated a view of frozen earth as an aggregate structure, ground. On the other hand, the process of expanding and asserting control over diverse and far-flung places gave rise to a “planetary consciousness” (Pratt 1992). It cultivated a habit of seeing connections in nature across geographical space and between local, regional, and global scales (Wulf 2015). This perspective nurtured a view of frozen earth as part of a cohesive planetary envelope defined by the presence and activity of ice. Although the utilitarian and earth systems approaches generated distinct views of frozen earth, both can trace their development to processes of colonialism.

The coloniality of the historiography of frozen earth sets up a final claim: it is necessary to recognize and critique how ideas about frozen earth have been shaped by colonialism and to recover and make room for other forms of understanding. Rather than expecting to escape from colonial legacies or purge them from the science of frozen earth, the goal is to foster critical awareness and distance. Such awareness and distance must extend to the term permafrost itself, which, despite its flaws, has persisted. Interrogating its implications and considering alternatives can clarify the interests, assumptions, and blind spots that people bring to frozen earth. Scholars must resist allowing a term and definition promulgated in the particular contexts of Soviet industrialization and US militarization to narrow the historiography of frozen earth. Historians, anthropologists, and geographers have begun to elucidate the multiple ontologies that humans have created for frozen earth as well as the heterogeneity and dynamism of the phenomenon itself. Exploring the richness of local interactions with frozen earth, past and present, may uncover narratives and metaphors that point to ways of living with frozen earth beyond aspirations of conquest and control.