Abstract
Earth's long‐term climate has been profoundly influenced by the episodic assembly and breakup of supercontinents at intervals of ca. 500 m.y. This reflects the cycle's impact on global sea level and atmospheric CO2 (and other greenhouse gases), the levels of which have fluctuated in response to variations in input from volcanism and removal (as carbonate) by the chemical weathering of silicate minerals. Supercontinent amalgamation tends to coincide with climatic cooling due to drawdown of atmospheric CO2 through enhanced weathering of the orogens of supercontinent assembly and a thermally uplifted supercontinent. Conversely, breakup tends to coincide with increased atmospheric CO2 and global warming as the dispersing continental fragments cool and subside, and weathering decreases as sea level rises. Supercontinents may also influence global climate through their causal connection to mantle plumes and large igneous provinces (LIPs) linked to their breakup. LIPs may amplify the warming trend of breakup by releasing greenhouse gases or may cause cooling and glaciation through sulfate aerosol release and drawdown of CO2 through the chemical weathering of LIP basalts. Hence, Earth's long‐term climatic trends likely reflect the cycle's influence on sea level, as evidenced by Pangea, whereas its influence on LIP volcanism may have orchestrated between Earth's various climatic states.
Keywords: atmospheric CO2 , climate, large igneous provinces, sea level, supercontinent cycle
The supercontinent cycle, by which Earth history is viewed as having been punctuated by the episodic assembly and breakup of supercontinents, has, through its management of plate motion, planetary geography, sea level and mantle circulation, profoundly influenced Earth's long‐term climatic history. It is likely that the supercontinent cycle has, over the course of Earth history, played a dominant role in governing the climate of individual continents, the planet's long‐term warming and cooling trends, and its occasional climatic extremes, while, at the same time, maintaining surface conditions sufficiently hospitable to ensure the continuity of life.
INTRODUCTION
The supercontinent cycle describes the realization, developed over the past 30 years, that much of Earth history has been punctuated by the episodic assembly and breakup of supercontinents, during which most of Earth's continents are assembled into a single landmass. 1 Consequently, the well‐documented supercontinent Pangea (Figure 1), first advocated by Wegener, 2 , 3 is viewed as only the most recent in a series of supercontinents that have assembled and broken up at intervals of roughly half‐a‐billion years since perhaps as far back as the late Archean. 4 , 5 , 6 , 7 , 8 Major support for this hypothesis has come with the recognition of supercontinents (Figure 2) at c. 620–580 Ma (Pannotia, 6 , 9 , 10 , 11 , 12 the existent of which is debated 13 , 14 , 15 ), c. 950–800 Ma (Rodinia 6 , 16 , 17 , 18 ), and c. 1.6–1.4 Ga (Nuna or Columbia 19 , 20 , 21 , 22 , 23 , 24 ), and possible supercontinents (or supercratons) at c. 2.7–2.5 Ga (Kenorland; 25 , 26 , 27 Lauroscandia 28 ) and c. 3.0 Ga (Ur 29 , 30 ), in addition to Wegener's Pangea (c. 325–200 Ma).
FIGURE 1.
Reconstruction of Pangea for the Late Triassic (at 200 Ma) by the PLATES program at the University of Texas Institute of Geophysics. (http://www‐udc.ig.utexas.edu/external/plates/images/pangea_07sep2007.jpg)
FIGURE 2.
Proposed reconstructions of pre‐Pangean supercontinents. (A) Pannotia (c. 620–580 Ma 238 ), (B) Rodinia (c. 950–800 Ma 239 simplified after Li et al. 17 ), (C) Nuna/Columbia (c. 1.6–1.4 Ga 23 ), and (D) Kenorland (c. 2.7–2.5 Ga). Abbreviations: (A) A, Australia; AM, Amazonia; AN, Antarctica; B, Baltica; BTS, Borborema–Trans‐Sahara; CSF, Congo–São Francisco; I, India; K, Kalahari; L, Laurentia; LP, Rio de la Plata; M, Madagascar; PA, Pampea; PR, Paraná; RA, Rio Apa; WA, West Africa. (B) A‐F, Albany‐Fraser orogen; EG, Eastern Ghats belt; K‐I, Kibaran and Irumide belts; M, Musgrave orogen; N‐N, Namaqua‐Natal province; S, Sunsas orogen; S‐N, Sveco‐Norwegian orogen; W, Wilkes province. (C) AM, Amazonia; BA, Baltica; CA, Cathaysia; EA, East Antarctica; LA, Laurentia; IN, India; NC, North China; NA, North Australia; SA, South Australia; SB, Siberia; WA, West Australia; WAF, West Africa. (D) Bund, Bundelkhand craton; NCC(EB), Eastern block of North China craton
The episodic cycle has been linked to global orogenesis, 31 , 32 , 33 granitoid magmatism and zircon age peaks, 34 , 35 , 36 , 37 crustal growth, 38 , 39 , 40 , 41 mineralization, 42 , 43 , 44 , 45 , 46 , 47 , 48 large igneous provinces (LIPs) 49 , 50 , 51 , 52 , 53 and deep mantle convection patterns. 54 , 55 , 56 , 57 , 58 , 59 , 60 Additionally, the cycle has been shown to have profound affects on sea level, 61 , 62 , 63 , 64 , 65 , 66 ocean chemistry, 35 , 67 , 68 , 69 the stable isotope record, 35 , 70 , 71 , 72 patterns of sedimentation, 73 , 74 , 75 atmospheric composition, 76 , 77 , 78 global biogeochemical cycles, 4 , 79 , 80 climate 74 , 81 , 82 , 83 , 84 marine biodiversity, 85 , 86 and the evolution of life. 83 , 87 , 88
The supercontinent cycle is consequently a unifying hypothesis with major implications for the geosciences and our understanding of Earth's evolution. It has likely influenced the rock record more than any other geologic phenomena, 89 its existence documents fundamental processes in the Earth's mantle and at the core–mantle boundary, 50 and it has probably governed the planet's surface environment for much of Earth history. 90
For detailed reviews of the history, development, and consequences of the supercontinent cycle, the reader is referred to Nance and Murphy 91 and Nance et al. 1 Here, I focus on just one aspect of the cycle—its affect on Earth's climate and climate‐controlling processes.
BACKGROUND
That Earth history may have been punctuated by the episodic assembly and breakup of supercontinents with profound consequences to the geosphere is not a new idea, 4 , 5 , 61 , 79 , 92 and the notion of long‐term episodicity in tectonic processes predates plate tectonics. 93 , 94 , 95 , 96 , 97 , 98 , 99 , 100 However, widespread recognition of the supercontinent cycle is a relatively recent phenomenon, 7 , 8 , 55 , 89 , 101 as is the growing consensus regarding its profound effect on Earth history and evolution. 40 , 80 , 81 , 83 , 90 , 102 , 103 , 104 , 105
A wide variety of phenomena have been linked to the supercontinent cycle (Figure 3). Supercontinent assembly, for example, is accompanied by terrane accretion, collisional orogenesis, and continental shortening as the continents amalgamate and the oceans between them close. 39 Orogenic granitoid magmatism, recorded as U–Pb age peaks for zircons with evolved εHf and elevated δ18O values consistent with increased reworking of crustal and sedimentary material, is enhanced, 34 , 35 , 41 , 106 , 107 as are conditions for continental arc magmatism, 108 extreme (UHT and UHP) metamorphism, 32 , 109 and active margin sedimentation with high clastic to carbonate ratios. 75 Epeirogenic uplift through continental insulation and mantle upwelling, both of which are thought to be consequences of supercontinent amalgamation, 56 , 110 , 111 lead to a global lowering of sea level 63 , 64 , 112 , 113 with accompanying enhanced weathering and terrestrial deposition. 76 The resulting drawdown of atmospheric CO2 causes climatic cooling, 114 , 115 while the loss of insular continents and shallow‐marine habitats leads to low biotic diversity 85 and may precipitate mass extinctions. Enhanced erosion increases seawater 87Sr/86Sr, δ34S and nutrient supply, 35 , 70 , 71 , 72 , 116 , 117 while the resulting rise in marine productivity and photosynthesis acts to increase atmospheric oxygen levels. 71 , 78
FIGURE 3.
Secular trends in detrital zircon ages, granulite facies thermal gradients, passive margin development, normalized seawater 87Sr/86Sr, and mean initial εHf and average δ18O in detrital zircons from recent sediments compared with the assembly (shaded intervals) of various proposed pre‐Pangean supercontinents (from Hawkesworth et al. 36 and references therein). Abbreviations: HP, high pressure; UHP, ultra‐high pressure; UHT, ultra‐high temperature
On the other hand, supercontinent breakup and dispersal reverses many of these trends and is heralded by peripheral subduction rollback 50 , 118 , 119 , 120 and continental rifting documented in mafic dike swarms and LIPs, 49 , 51 , 53 , 121 followed by passive margin development. 75 , 122 Subdued collisional orogeny and granitoid magmatism is recorded in troughs in U–Pb age spectra for zircons with juvenile εHf and lowered δ18O values consistent with increased mantle‐derived magmatism. 34 , 35 , 41 , 107 Thermal subsidence and extension of the dispersing continental fragments, and the creation of a younger world ocean floor through the opening of new ocean basins and consequent increase in ridge length, is accompanied by rapid sea level rise, 63 , 64 , 66 , 123 enhanced shallow marine sedimentation, 90 and organic carbon burial leading to negative δ13C anomalies. 70 , 83 Diminished seawater 87Sr/86Sr ratios and warm, equable climates are linked to elevated atmospheric CO2 levels, driving rapid evolutionary radiation of new taxa and increasing biotic diversity. 35 , 70 , 71 , 78 , 124
The cycle is likely driven by some combination of continental insulation, mantle plume dynamics, and slab rollback. The mechanism first proposed was that of continental insulation, 61 , 110 , 125 whereby the thermal insulating effect of continental lithosphere on mantle heat flow is considered to trap mantle heat beneath supercontinents resulting in their thermal uplift and breakup. 56 , 111 , 126 The new oceans so produced then either widen until the leading edges of the dispersing continental fragments collide to form a new supercontinent, a process termed extroversion, 127 or they close as their floors grow older and less buoyant, such that the continental fragments are reassembled, a process termed introversion. In both cases, the assembled supercontinent would once again trap mantle heat and the cycle would be repeated.
Alternatively, the mechanism may be a consequence of the cycle's strong coupling to mantle dynamics, 60 whereby subduction to the core–mantle boundary of the oceanic lithosphere surrounding a supercontinent creates mantle plumes that rise beneath them and contribute to their breakup. 38 , 50 , 54 , 128 In this case (Figure 4), supercontinents are considered to form over areas of mantle downwelling in an Earth with a degree‐1 mantle structure, that is, one with single, antipodal zones of mantle upwelling and downwelling. 54 They subsequently break up because the subduction girdle that develops around a supercontinent once it assembles creates a slab graveyard of subducted oceanic lithosphere at the core–mantle boundary, 129 , 130 which influences the mantle's large low shear velocity provinces (LLSVPs) in such a way as to foster the generation of mantle plumes that rise beneath the supercontinent. 50 , 59 , 60 , 131 The result is an Earth with a degree‐2 mantle structure, that is, one with two antipodal zones of upwelling, the one beneath the supercontinent being responsible for its breakup. 50 , 54 Upon breakup, the subduction girdle that develops around the supercontinent following its assembly forms a new ring of mantle downwelling over which the dispersing continental fragments gather. This girdle, which would be longitudinal if true polar wander brings a supercontinent to the equator, 50 , 54 , 132 may then move away from the former supercontinent to recreate an antipodal degree‐1 mantle structure and reassemble a supercontinent by way of extroversion, or it may move toward the former supercontinent and reassemble one by way of introversion. 127 Alternatively, the dispersing continental fragments may coalesce along the girdle such that the new supercontinent assembles roughly 90 degrees away from its predecessor, a process termed orthoversion. 133
FIGURE 4.
Numerical modeling of supercontinent assembly and breakup. 50 (A) Initial small‐scale convection evolves to (B) an early stage degree‐1 mantle structure (antipodal regions of upwelling and downwelling) as the supercontinent assembles, and (C) a stable degree‐1 structure as the supercontinent forms. (D) With the formation of a subduction girdle and the onset of a superplume beneath the supercontinent, convection evolves to a degree‐2 planform (antipodal regions of upwelling), which (E) contributes to supercontinent breakup as true polar wander brings the supercontinent to the equator. Alternation of the two modes of mantle convection is thought to be responsible for the cyclic process of supercontinent assembly and breakup. Blue = cool mantle, yellow = hot mantle, red = core
A potential breakup mechanism also exists in the forces associated with slab rollback along the supercontinent periphery. 118 , 120 , 134 , 135 , 136 This mechanism is consistent with the development of a slab girdle, the oceanward retreat of which would generate extensional forces that may be sufficient to cause supercontinent breakup. 118
All three mechanisms are supported by modeling, 55 , 118 , 125 , 126 and it is likely that each plays a role in the breakup of supercontinents once they have amalgamated. Hence, the cycle appears to operate because supercontinents sow the seeds of their own destruction and break up, but in doing so, they set the stage for their eventual reassembly. While their relationship to the supercontinent cycle is unlikely to be a simple one, 52 , 137 , 138 the apparent role of mantle plumes is significant because it links the supercontinent cycle to deep mantle upwelling and processes occurring at the core–mantle boundary. Hence, it elevates the supercontinent cycle from a near‐surface phenomena to a whole‐mantle process linking top‐down plate tectonics and bottom‐up plume tectonics.
INFLUENCE ON GLOBAL CLIMATE
The role of the supercontinent cycle in governing long‐term global climate is chiefly based on the Phanerozoic record and rests largely on its influence on global sea level and the governing affect this has on continental erosion and silicate weathering, and the consequent abundance of CO2 and other greenhouse gases in the atmosphere. 4 , 76 , 139 , 140 , 141 However, the cycle also influences climate through its control of continental geography and through the association of supercontinent amalgamation and breakup with LIP events. 49 , 52 , 142 LIP events have been correlated with a wide variety of environmental impacts and can profoundly influence global climate, both through the release of large volumes of volcanic CO2 to the atmosphere 143 , 144 and through extreme atmospheric CO2 drawdown brought about by the weathering of equatorial flood basalts. 145
Influence on global sea level
The supercontinent cycle has a profound effect on global sea level as a result of its long‐term control of both the elevation of the continents and the depth of the ocean basins. 62 , 63 , 64 , 66 In fact, the close correspondence between the changes in global sea level predicted by the cycle for the Phanerozoic, 61 which amounted to several hundred meters, and the contemporary depositional record of sea level change over the same interval 146 was a key argument used in support of the original hypothesis (Figure 5). Supercontinents tend to correspond to intervals of very low global sea level 112 , 147 as a result of their epeirogenic uplift, either because continental insulation traps mantle heat beneath them, and/or because descent of the subduction girdle to the core–mantle boundary fosters mantle upwelling beneath them. Shortening of the crust as a result of the collisional orogenies of supercontinent assembly may also lower sea level by increasing oceanic area. 61
FIGURE 5.
Comparison of the effect of the supercontinent cycle on sea level (straight‐segmented line), calculated for the Phanerozoic 61 given the known duration of Pangea (box), with the first‐order eustasy curve (undulating line). 146 The close correspondence between these two lines was used by Worsley et al. 61 to support their case for a supercontinent cycle. 91
Conversely, supercontinent breakup tends to correspond to a rapid global rise in sea level as a combined result of the thermal subsidence of the continental fragments as they disperse and cool, crustal extension as a result of rifting, and the decrease in ocean basin volume caused by the overall decrease in seafloor age and increase in the volume of mid‐ocean ridges that accompany the opening of new ocean basins floored by young oceanic lithosphere. 61 This rise in sea level results in widespread continental flooding, but is ultimately reversed as the new ocean basins get older.
Influence on atmospheric composition
Because of its demonstrated effect on Phanerozoic global sea level, the supercontinent cycle has likely had a profound influence on the long‐term levels of CO2 (and other greenhouse gases) in the atmosphere (Figure 6). Atmospheric CO2 levels have fluctuated throughout much of Earth history in response to variations in the input of this gas from volcanic exhalations and the breakdown of carbonates and organic matter, and its removal through the chemical weathering of the continents and photosynthesis, 148 , 149 , 150 the former involving its reaction with Ca and Mg silicates to form Ca and Mg carbonates following riverine transport of the weathering products to the oceans. 151 Since the efficacy of this process depends, in part, on the land area available for chemical weathering, its effect on atmospheric CO2 levels, and hence climate, varies with sea level. Hydrothermal alteration of seafloor basalts likely provides an independent sink for atmospheric CO2, 152 , 153 , 154 while the subduction of platform carbonates at continental margin arcs may provide a significant additional source. 155
FIGURE 6.
Phanerozoic proxy reconstructions and modeled predictions (Geocarb III 150 ) of atmospheric CO2 levels for the Phanerozoic. 241 Shaded area represents error range in modeling.
Supercontinent amalgamation and breakup
As a consequence of the relationship between land area and atmospheric CO2, supercontinents tend to coincide with climatic cooling due to atmospheric CO2 drawdown because they are associated with very low sea levels as a result of their thermal uplift. Adding to this cooling influence is the enhanced chemical weathering of the orogens of supercontinent assembly. Both of these processes would be amplified if true polar wander brings the supercontinent to the equator as a consequence of centrifugal forces acting on the positive dynamic topography (excess mass) created by its thermal uplift, 50 , 54 , 132 since the reaction rates of chemical weathering, and hence the rate of drawdown of atmospheric CO2, are strongly dependent on temperature and precipitation. 76 , 156
As a likely result of these processes, the amalgamation of both Pangea and Pannotia was accompanied by cold, “icehouse” climates (Figure 7) and widespread continental glaciation—respectively, the c. 335–260 Ma Gondwanan 115 , 157 , 158 and the c. 640–635 Ma Marinoan 159 and c. 580/565 Ma Gaskiers/post‐Gaskiers. 160 , 161 , 162 , 163 Conversely, continental glaciation accompanied the breakup of Rodinia (c. 717–663 Ma Sturtian 164 , 165 , 166 ), Kenorland/Lauroscandia (c. 2.44–2.3 Ga Huronian [Gowganda] 167 , 168 , 169 ), and perhaps even the earliest proposed supercraton Ur (c. 2.9 Ga Pongola 170 , 171 , 172 ). This could reflect the abrupt erosional release of dissolved Ca and Mg to the oceans following the onset of rifting, 145 , 173 the combination of uplift and subsidence in rift settings having been long thought to provide ideal conditions for both the initiation of glaciation and the preservation of the resulting glacigenic sediments. 174 , 175 The role of the supercontinent cycle in continental glaciation, however, is a complex one, and while supercontinents may foster ice ages, they do not mandate them as evidenced by the apparent absence of any glaciation associated with Nuna/Columbia and its unrelated presence during the Hirnantian (c. 445 Ma) 176 , 177 and the Pleistocene to present day.
FIGURE 7.
Distribution of warm (greenhouse) and cool (icehouse) global climatic conditions for the past 1 Ga 124 compared with times of supercontinent assembly and breakup for Rodinia, Pannotia, and Pangea.
The Huronian glaciations accompanying Kenorland/Lauroscandia also coincide with the Great Oxidation Event (c. 2.43–2.25 Ga 178 ), during which biologically produced O2 first started to accumulate in the atmosphere, 179 perhaps as a result of the breakup‐related evolution of the first oxygen‐requiring cyanobacteria, 180 or a LIP‐generated pulse of sulphate to the oceans, the reduction of which liberated oxygen. 181 The rise in atmospheric oxygen, evident in the loss of Fe‐poor paleosols, detrital pyrite, and detrital uraninite, 182 , 183 , 184 in the first appearance of redbeds, 185 and in the loss of mass‐independent fractionation of sulfur isotopes in sedimentary rocks, 186 , 187 likely led to the demise of atmospheric methane, the most powerful of the greenhouse gases, thereby providing an alternative mechanism for dramatic climatic cooling. 188 , 189 , 190
Snowball Earth
The climatic cooling that led to the continental glaciations associated with Kenorland or Lauroscandia (Huronian/Gowganda), Rodinia (Sturtian), and Pannotia (Marinoan) is thought to have been sufficiently extreme as to cause the entire planet to freeze, a unique situation known as “Snowball Earth.” 81 , 164 , 189 , 190 , 191 , 192 Such conditions are thought possible if ice comes to within c. 30o of the equator because the albedo feedback from the planet's ice‐covered surface then becomes self‐sustaining 193 , 194 , 195 —one more latitudinal degree of ice cover causing albedo cooling sufficient to give one more latitudinal degree of cover (Figure 8). As a result, glacial ice spreads rapidly toward the equator, eventually leading to an ice‐covered planet with a global mean temperature estimated at c. −50°C. 192 In the case of the Sturtian (c. 717–663 Ma 165 , 196 ) and Marinoan (c. 640–635 Ma 159 ) glaciations, such Snowball Earth conditions were likely promoted by the concentration of continents between 30oN and 30oS, and the consequent high rates of chemical weathering and atmospheric CO2 drawdown, following the breakup of Rodinia, 145 the final equatorial position of which 17 may have been the result of true polar wander. 54
FIGURE 8.
Time scale for estimated changes in global mean surface temperature, based on energy‐balance calculations, and ice extent through one complete snowball event. 192 The global palaeogeography is for 750 Ma, some 30 m.y. before the Sturtian glaciation. Abbreviations: Am, Amazonia; Au, Australia; Ba, Baltica; Co, Congo; In, India; K, Kalahari; M, Mawson; Si, Siberia; Ta, Tarim; WA, West Africa; Y, South China
Once initiated, the icehouse conditions of a Snowball Earth are thought to prevail until volcanically sourced atmospheric CO2, deprived by ice cover of a continental weathering (and photosynthetic) sink, rises dramatically to c. 350 present atmospheric levels. 81 , 197 At this threshold point, rapid greenhouse‐induced and albedo‐feedback accelerated deglaciation ensues (Figure 8), leading within c. 5 Ma, to a “hothouse” Earth with a global mean temperature of c. 40°C. 192 With re‐establishment of the CO2 cycle and renewal of continental weathering (and photosynthesis), the climate rapidly returns to its initial state, setting the stage for the process to repeat. There are consequently five stages in the evolution of a Snowball Earth: (1) strong equatorial drawdown of atmospheric CO2 through continental weathering needed to cause the oceans to start freezing, (2) albedo‐feedback expansion of the ice cover to a latitude of c. 30o, whereupon it becomes self‐sustaining and the planet freezes from pole to pole, (3) shutdown of continental weathering allowing volcanically derived CO2 to build rapidly in the atmosphere, (4) greenhouse effect of rising atmospheric CO2 levels reaches a critical threshold, whereupon the ice rapidly melts and a hothouse world is established, and (5) resumption of continental weathering and restoration of the CO2 cycle reduces the greenhouse effect and returns climate to its initial state.
Supercontinent dispersal
The processes that lead to global cooling during the assembly and rifting of supercontinents are reversed following supercontinent breakup as the dispersing continental fragments cool and subside. With the ensuing rise in global sea level, the continents flood, continental weathering decreases, and atmospheric CO2 levels rise. As a result, continental dispersal tends to coincide with a progressive build‐up of atmospheric CO2 and accompanying global warming. In addition, the release of CH4 (or the CO2 produced by its oxidation) as gas hydrates break down with rising temperatures would provide this breakup‐related global warming with a strong positive feedback. 198 , 199 Not surprisingly, therefore, supercontinent breakup tends to coincide with climatic warming, as evidenced by the “greenhouse” climates of the Mesozoic, early Paleozoic, and much of the Tonian, 124 , 200 , 201 following the breakup of Pangea, Pannotia, and Rodinia, respectively (Figure 7). The introduction of large amounts of CO2 into the oceans during supercontinent breakup and dispersal has also been linked to increased carbon burial and black shale abundance, 70 while the increased run‐off of terrigenous nutrients in warmer climates has been coupled to oceanic anoxia. 202 , 203
An additional climatic influence of supercontinent breakup comes from its proposed link to stepwise increases in atmospheric oxygen, possibly as a consequence of enhanced marine productivity resulting from an increase in the erosional release to the oceans of nutrients, such as bioproductivity‐limiting phosphorus. 78 Like CO2, atmospheric O2 levels are thought to have had a significant impact on long‐term global climate, 204 even though oxygen is not a greenhouse gas. This is because rising O2 levels result in an increase in atmospheric density and, hence, greater scattering of incoming solar radiation and consequent reduction in surface evaporation. As a result, precipitation decreases, humidity levels fall, and cooler temperatures ensue because less heat is trapped by water vapor, which is a strong greenhouse gas.
Increased atmospheric O2 levels might also be expected during periods of enhanced organic carbon burial, such as those proposed to accompany the rapid sedimentation of supercontinent breakup and dispersal. 78 , 117 Conversely, decreased atmospheric O2 levels should accompany the increased chemical weathering of supercontinent amalgamation and breakup because the chemical reactions involved are largely oxidative. 204
Influence on mantle plumes and LIPs
Since supercontinent breakup requires continents to rift, the supercontinent cycle has long been linked to mafic dike swarms and LIPs, 4 , 49 , 61 , 205 , 206 and through their emplacement, to the activity of mantle plumes. 52 , 58 , 60 , 207 , 208 Uncertainty continues to exist as to whether the timing of LIP events (Figure 9) coincides with the breakup of supercontinents, 4 , 61 , 205 , 209 or their amalgamation, 49 , 206 or both, 137 , 138 , 210 in part because the timing and number of pre‐Pangean supercontinent amalgamation and breakup events remain poorly constrained 211 even while the dating of LIP events has become increasingly precise. 51 , 53 , 142 However, while evidence has been presented that questions the relationship, 138 , 212 , 213 recent time‐series analysis suggests a cyclicity in both continental and oceanic LIPs and accompanying plume activity that is both comparable to that of the supercontinent cycle and corresponds closely to periods of supercontinent rifting and breakup. 60 This is consistent with the idea that supercontinent amalgamation works to trigger mantle plumes at the core–mantle boundary; 50 , 128 a proposition that finds support in the correlation between the reconstructed positions of Mesozoic LIPs and the margins of the African (Tuzo) LLSVP, which has been identified at the core–mantle boundary on the basis of seismic tomography, and which is centered over the former position of Pangea. 214 , 215 , 216
FIGURE 9.
Distribution of large igneous provinces (LIPs) throughout Earth history 137 compared with tenure of supercontinents/supercratons Pangea (c. 325–200 Ma), Pannotia (c. 620–580 Ma), Rodinia (c. 950–800 Ma), Nuna/Columbia (1.6–1.4 Ga), Kenorland (c. 2.7–2.5 Ga), and Ur (c. 3 Ga)
LIPs and climate
Mantle plumes can, in and of themselves, affect climate simply by thermally uplifting the lithosphere and thereby changing global sea level and weathering‐mediated atmospheric CO2 levels. 217 , 218 The influence of LIPs on global climate, however, stems from the voluminous volcanic activity with which they are associated, the effect of which can cause both climatic warming and cooling. The immediate effect of this volcanism is one of brief regional or global cooling as a result of the dispersal and absorption of solar radiation by fine volcanic ash and H2SO4 aerosols vented to the stratosphere during explosive eruptions. 219 However, the most dramatic climatic effect of LIPs is one of long‐term global warming due to the increased magmatic venting of greenhouse gases, such as CO2 and CH4. 144 The introduction of such gases to the atmosphere during supercontinent rifting and breakup would act to boost those generated by the decrease in continental weathering associated with breakup‐related sea level rise, further enhancing global warming. Depending on the rock‐type, contact metamorphism associated with LIP magmatism can also release huge volumes of greenhouse gases to the atmosphere. 220 In fact, these may play a leading role in global warming, given that the dominant LIP magma is relatively gas‐poor tholeiitic basalt.
In addition to their climatic impact, LIP magmatism and contact metamorphism liberate large volumes of toxic gases, such as SO2 and F, so it is not surprising that Phanerozoic LIP flood volcanism has long been correlated with mass extinctions (Figure 10). 88 , 144 , 221 , 222 A strong correlation exists, for example, between the Yakutsk‐Vilyui, 223 , 224 Emeishan, 225 Siberian Traps, 226 , 227 CAMP, 228 Karoo‐Ferrar, 229 and Deccan Traps 230 LIP events and mass extinctions in the Late Devonian (Frasnian‐Famennian), Middle Permian (Capitanian), end‐Permian, end‐Triassic, Early Jurassic (Toarcian), and end‐Cretaceous, respectively. 231 A temporal link also exists between the final pulses of the Central Iapetus Magmatic Province (CIMP) 143 and the extinction of the Ediacaran fauna immediately prior to the Cambrian explosion. 232
FIGURE 10.
Age and estimated volume of Phanerozoic large igneous provinces (LIPs) compared to genus extinction magnitude showing correlation between mass extinction events (peaks) and LIP emplacement, particularly during tenure of Pangea. 88 Large igneous provinces: PDD, Pripyat‐Dnieper‐Donets; CAMP, Central Atlantic Magmatic Province; OJP 1/OJP 2, Ontong Java Plateau phases 1 and 2; NAIP, North Atlantic Igneous Province; CR, Columbia River Basalt Group. Extinction events: EMC, Early to Middle Cambrian; IME, Ireviken, Mulde and Lau Events; I‐D, intra‐Devonian events; F/F, Frasnian/Famennian; D/C, Devonian/Carboniferous; P/T, Permian/Triassic; T/J, Triassic/Jurassic; K/ Pg, Cretaceous/Paleogene. Stars identify bolide impacts.
However, the long‐term warming influence of major LIP events may be followed, or interrupted, by abrupt cooling. It has been argued, for example, that the equatorial continental paleogeography of the Cryogenian, which would have favored cool global climates as a result of climate‐enhanced chemical weathering and organic carbon burial, 233 may have been driven into runaway global glaciation of the Sturtian Snowball Earth by the weathering of extensive LIP continental flood basalts erupted throughout the break‐up of Rodinia, such as those associated with the Gunbarrel (c. 780 Ma), Mundine Well (c. 755 Ma), and Franklin (c. 723 Ma) provinces. 145 , 234 , 235 Donnadieu et al. 145 further suggest that this may also have been the case for the Marinoan Snowball Earth.
Other factors
According to Jellinek et al., 84 an additional influence of the supercontinent cycle on global climate may lie in its control on the degree to which warm subcontinental mantle is globally mixed, since the impact of volcanism and weathering on Earth's long‐term carbon cycle is modulated by lateral ocean‐continent variations in mantle temperature. Their calculations suggest that supercontinents girdled by subduction zones foster lateral ocean‐continent mantle temperature variations because mixing of insulated subcontinental mantle is inhibited. As a result, outgassing of CO2 from mid‐ocean ridges is reduced, giving rise to cold climates and icehouse/hothouse climate variability like that associated with Rodinia. Conversely, long‐lived ice‐free climates, like that associated with Nuna/Columbia, are features of thorough mantle thermal mixing.
The supercontinent cycle can also influence climate solely as a result of the changes it makes in the distribution of continents and oceans. By applying a climate system model to the breakup of Pangea, for example, Tabor et al. 236 have shown that opening of an ocean basin such as the Atlantic fosters humidification of the tropics, large‐scale reorganization of tropical circulation, and both regional and global changes in temperature. Weaker tropical easterlies and reduced upwelling warm the equatorial ocean, while increased moisture and cloud formation in the tropics cool both land and sea.
Finally, as pointed out by Foley and Driscoll, 237 plate tectonics, as governed by the supercontinent cycle, is itself influenced by climate. Cool climates, which the cycle maintains through its plate tectonic control of the long‐term carbon cycle, act to enhance stresses within the lithosphere and promote its hydration and weakening that, in turn, enable plate tectonics to take place. Hence, the supercontinent cycle may have played a significant role in ensuring Earth maintained its status as a habitable planet.
CONCLUSIONS
The supercontinent cycle, by which Earth history is viewed as having been punctuated by the episodic assembly and breakup of supercontinents, has, through its management of plate motion, planetary geography, sea level, and mantle circulation, profoundly influenced Earth's long‐term climatic history. By necessitating alternating episodes of supercontinent assembly, during which the continents approach one another, and breakup, during which they disperse, the cycle has governed Earth's paleogeography and, in doing so, the regional climate experienced by any given continent at any given time. 69 By exercising control over the drawdown of CO2 and other greenhouse gases from the atmosphere through its influence on sea level and chemical weathering, and the input of these gases to the atmosphere through its influence on plate tectonics and magmatism, the cycle has mediated Earth's long‐term global record of alternating warm (greenhouse) and cold (icehouse) climates. A strong coupling also appears to exist between supercontinents and mantle dynamics that would link the cycle to mantle plumes and LIPs, and, consequently, the climatic effects of their volcanic emissions, which have been associated with mass extinctions, oceanic anoxia, and catastrophic changes to the surface environment. The proposed tendency for true polar wander to center supercontinents on the equator as a result of centrifugal forces acting on their excess mass may also set the stage for extreme global cooling (Snowball Earth) through the enhanced drawdown of atmospheric CO2 caused by the equatorial weathering of breakup‐related LIP basalts. It is, therefore, likely that the supercontinent cycle has, over the course of Earth history, played a dominant role in governing the climate of individual continents, the planet's long‐term warming and cooling trends, and its occasional climatic extremes, while, at the same time, maintaining surface conditions sufficiently hospitable to ensure the continuity of life.
COMPETING INTERESTS
The author declares no competing interests.
PEER REVIEW
The peer review history for this article is available at https://publons.com/publon/10.1111/nyas.14849.
ACKNOWLEDGMENTS
This review has benefited from the insightful comments of Ian Dalziel, Peter Cawood, and Sergei Pisarevsky, whose appraisals of the paper were greatly appreciated. The paper has also benefited from the comments of Kent Condie and Daniel Pastor‐Galán, whose reviews of an earlier version of the manuscript proved equally useful. The review was completed during a visiting professorship at Charles University in Prague funded by the Czech Operational Programme Research, Development and Education, Project no. CZ.02.2.69/0.0/0.0/16_015/0002362. This support and the assistance and fellowship of Jiří Žák, Jarka Hajná, and all their colleagues at the Institute of Geology and Paleontology are gratefully acknowledged.
Nance, R. D. (2022). The supercontinent cycle and Earth's long‐term climate. Ann NY Acad Sci., 1515, 33–49. 10.1111/nyas.14849
REFERENCES
- 1. Nance, R. D. , Murphy, J. B. , & Santosh, M. (2014). The supercontinent cycle: A retrospective essay. Gondwana Research, 25, 4–29. [Google Scholar]
- 2. Wegener, A. (1915). Die Entstehung der Kontinente und Ozeane . Sammlung Vieweg (Vol. 23). Braunschweig: Druck and von Freidrich Vieweg. [Google Scholar]
- 3. Wegener, A. (1920). Die Entstehung der Kontinente und Ozeane . Die Wissenschaft Band 66. (2nd ed.). Braunschweig: Druck and von Freidrich Vieweg. [Google Scholar]
- 4. Worsley, T. R. , Moody, J. B. , & Nance, R. D. (1985). Proterozoic to recent tectonic tuning of biogeochemical cycles. In Sundquist, E. T. , & Broecker, W. S. (Eds.), The carbon cycle and atmospheric CO2: Natural variations, Archean to present (pp. 561–572). American Geophysical Union. [Google Scholar]
- 5. Worsley, T. R. , Nance, R. D. , & Moody, J. B. (1986). Tectonic cycles and the history of the earth's biogeochemical and paleoceanographic record. Paleoceanography, 1, 233–263. [Google Scholar]
- 6. Dalziel, I. W. D. (1997). Neoproterozoic–Paleozoic geography and tectonics: Review, hypothesis, environmental speculation. Geological Society of America Bulletin, 108, 16–42. [Google Scholar]
- 7. Rogers, J. J. W. , & Santosh, M. (2003). Supercontinents in Earth history. Gondwana Research, 6, 357–368. [Google Scholar]
- 8. Evans, D. A. D. (2013). Reconstructing pre‐Pangean supercontinents. Geological Society of America Bulletin, 125, 1735–1751. [Google Scholar]
- 9. Stump, E. (1987). Construction of the Pacific margin of Gondwanaland during the Pannotios cycle. In Mckenzie, G. D. (Ed.), Gondwana six: Structure, tectonics and geophysics (pp. 77–87). American Geophysical Union. [Google Scholar]
- 10. Stump, E. (1992). The Ross orogen of the Transantarctic Mountains in the light of the Laurentian–Gondwana split. GSA Today, 2, 25–27. [Google Scholar]
- 11. Powell, C. M. C. A. (1995). Are Neoproterozoic glacial deposits preserved on the margins of Laurentia related to the fragmentation of two supercontinents? [Comment]. Geology, 23, 1053–1054. [Google Scholar]
- 12. Dalziel, I. W. D. (2013). Antarctica and supercontinental evolution: Clues and puzzles. Earth and Environmental Science Transactions of the Royal Society of Edinburgh, 104, 3–16. [Google Scholar]
- 13. Nance, R. D. , & Murphy, J. B. (2019). Supercontinents and the case for Pannotia, In Wilson, R. W. , Houseman, G. A. , McCaffrey, K. J. W. , Doré, A. G. , & Buiter, S. J. H. (Eds.), Fifty Years of the Wilson Cycle Concept in Plate Tectonics (pp. 65–85), Geological Society of London, Special Publications. [Google Scholar]
- 14. Evans, D. A. D. (2021). Pannotia under prosection. In Murphy, J. B., Strachan, R. A., & Quesada, C. (Eds.), Pannotia to Pangaea: Neoproterozoic and Paleozoic orogenic cycles in the Circum‐Atlantic Region (pp. 63–81). Geological Society, London, Special Publications. [Google Scholar]
- 15. Nance, R. D. , Evans, D. A. D. , & Murphy, J. B. (2023). Pannotia: To be or not to be. In Scotese, C. , Muller, D. , & van Hinsbergen, D. J. J. (Eds.), Plate tectonics, the last 2 billion years: Foundations of the earth system. Earth‐Science Reviews. In press. [Google Scholar]
- 16. McMenamin, M. A. S. , & McMenamin, D. L. S. (1990). The emergence of animals: The Cambrian breakthrough. New York: Columbia University Press. [Google Scholar]
- 17. Torsvik, T. H. (2003). The Rodinia jigsaw puzzle. Science, 300, 1379–1381. [DOI] [PubMed] [Google Scholar]
- 18. Li, Z. X. , Bogdanova, S. V. , Collins, A. S. , Davidson, A. , De Waele, B. , Ernst, R. E. , Fitzsimons, I. C. W. , Fuck, R. A. , Gladkochub, D. P. , Jacobs, J. , Karlstrom, K. E. , Lu, S. , Natapov, L. M. , Pease, V. , Pisarevsky, S. A. , Thrane, K. , & Vernikovsky, V. (2008). Assembly, configuration, and break‐up history of Rodinia: A synthesis. Precambrian Research, 160, 179–210. [Google Scholar]
- 19. Hoffman, P. F. (1997). Tectonic genealogy of North America. In Van der Pluijm, B. A. , & Marshak, S. (Eds.), Earth structure: An introduction to structural geology and tectonics (pp. 459–464). New York: McGraw‐Hill. [Google Scholar]
- 20. Rogers, J. J. W. , & Santosh, M. (2002). Configuration of Columbia, a Mesoproterozoic supercontinent. Gondwana Research, 5, 5–22. [Google Scholar]
- 21. Zhao, G. , Cawood, P. A. , Wilde, S. A. , & Sun, M. (2002). Review of global 2.1–1.8 Ga collisional orogens and accreted cratons: A pre‐Rodinia supercontinent? Earth‐Science Reviews, 59, 125–162. [Google Scholar]
- 22. Zhao, G. , Sun, M. , Wilde, S. A. , & Li, S. (2004). A Paleo‐Mesoproterozoic supercontinent: Assembly, growth and breakup. Earth‐Science Reviews, 67, 91–123. [Google Scholar]
- 23. Zhang, S. , Li, Z.‐X. , Evans, D. A. D. , Wu, H. , Li, H. , & Dong, J. (2012). Pre‐Rodinia supercontinent Nuna shaping up: A global synthesis with new paleomagnetic results from North China. Earth and Planetary Science Letters, 353‐354, 145–155. [Google Scholar]
- 24. Meert, J. G. , & Santosh, M. (2017). The Columbia supercontinent revisited. Gondwana Research, 50, 67–83. [Google Scholar]
- 25. Williams, H. , Hoffman, P. F. , Lewry, J. F. , Monger, J. W. H. , & Rivers, T. (1991). Anatomy of North America: Thematic portrayals of the continent. Tectonophysics, 187, 117–134. [Google Scholar]
- 26. Aspler, L. B. , & Chiarenzelli, J. R. (1998). Two Neoarchean supercontinents? Evidence from the Paleoproterozoic. Sedimentary Geology, 120, 75–104. [Google Scholar]
- 27. Lubnina, N. V. , & Slabunov, A. I. (2011). Reconstruction of the Kenorland supercontinent in the Neoarchean based on paleomagnetic and geological data. Moscow University Geology Bulletin, 66, 242. [Google Scholar]
- 28. Mints, M. V. , & Eriksson, P. G. (2016). Secular changes in relationships between plate‐tectonic and mantle‐plume engendered processes during Precambrian time. Geodynamics and Tectonophysics, 7, 173–232. [Google Scholar]
- 29. Rogers, J. J. W. (1996). A history of continents in the past three billion years. Journal of Geology, 104, 91–107. [Google Scholar]
- 30. Eriksson, P. G. , Banerjee, S. , Nelson, D. R. , Rigby, M. J. , Catuneanu, O. , Sarkar, S. , Roberts, R. J. , Ruban, D. , Mtimkulu, M. N. , & Sunder Raju, P. V. (2009). A Kaapval craton debate: Nucleus of an early small supercontinent or affected by an enhanced accretion event? Gondwana Research, 15, 354–372. [Google Scholar]
- 31. Nance, R. D. , & Murphy, J. B. (1994). Orogenic style and configuration of supercontinents. In Embry, A. F. , Beauchamp, B. , & Glass, D. J. (Eds.), Pangea: Global environments and resources (pp. 49–65). Canadian Society of Petroleum Geologists. [Google Scholar]
- 32. Brown, M. (2007). Metamorphism, plate tectonics, and the supercontinent cycle. Earth Science Frontiers, 14, 1–18. [Google Scholar]
- 33. Cawood, P. A. , Strachan, R. A. , Pisarevsky, S. A. , Gladkochub, D. P. , & Murphy, J. B. (2016). Linking collisional and accretionary orogens during Rodinia assembly and breakup: Implications for models of supercontinent cycles. Earth and Planetary Science Letters, 449, 118–126. [Google Scholar]
- 34. Condie, K. C. , Belousova, E. , Griffin, W. L. , & Sircombe, K. N. (2009). Granitoid events in space and time: Constraints from igneous and detrital zircon age spectra. Gondwana Research, 15, 228–242. [Google Scholar]
- 35. Condie, K. C. , & Aster, R. C. (2013). Refinement of the supercontinent cycle with Hf, Nd and Sr isotopes. Geoscience Frontiers, 4, 667–680. [Google Scholar]
- 36. Hawkesworth, C. J. , Cawood, P. A. , & Dhuime, B. (2016). Tectonics and crustal evolution. GSA Today, 26, 4–11. [Google Scholar]
- 37. Domeier, M. , Magni, V. , Hounslow, M. W. , & Torsvik, T. H. (2018). Episodic zircon age spectra mimic fluctuations in subduction. Scientific Reports, 8, 17471. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 38. Roberts, N. M. W. (1998). Episodic continental growth and supercontinents: A mantle avalanche connection? Earth and Planetary Science Letters, 21, 994–1000. [Google Scholar]
- 39. Condie, K. C. , & Aster, R. C. (2010). Episodic zircon age spectra of orogenic granitoids: The supercontinent connection and continental growth. Precambrian Research, 180, 227–236. [Google Scholar]
- 40. Hawkesworth, C. , Cawood, P. , & Dhuime, B. (2013). Continental growth and the crustal record. Tectonophysics, 609, 651–660. [Google Scholar]
- 41. Van Kranendonk, M. J. , & Kirkland, C. L. (2016). Conditioned duality of the Earth system: Geochemical tracing of the supercontinent cycle through Earth history. Earth‐Science Reviews, 160, 171–187. [Google Scholar]
- 42. Barley, M. E. , & Groves, D. I. (1992). Supercontinent cycles and the distribution of metal deposits through time. Geology, 20, 291–294. [Google Scholar]
- 43. Groves, D. I. (2005). Secular changes in global tectonic processes and their influence on the temporal distribution of gold‐bearing mineral deposits. Economic Geology, 100, 203–224. [Google Scholar]
- 44. Cawood, P. A. , & Hawkesworth, C. J. (2013). Temporal relations between mineral deposits and global tectonic cycles. In Jenkin, G. R. T. , Lusty, P. A. J. , McDonald, I. , Smith, M. P. , Boyce, A. J. , & Wilkinson, J. J. (Eds.), Ore deposits in an evolving Earth (pp. 9–21). Geological Society, London, Special Publications. [Google Scholar]
- 45. Hazen, R. M. , Liu, X.‐M. , Downs, R. T. , Golden, J. , Pires, A. J. , Grew, E. S. , Hystad, G. , Estrada, C. , & Sverjensky, D. A. (2014). Mineral evolution: Episodic metallogenesis, the supercontinent cycle, and the coevolving geosphere and biosphere. Society of Economic Geologists, 18, 1–15. [Google Scholar]
- 46. Bradley, D. C. (2015). Mineral evolution and Earth history. American Mineralogist, 100, 4–5. [Google Scholar]
- 47. Pirajno, F. , & Santosh, M. (2015). Mantle plumes, supercontinents, intracontinental rifting and mineral systems. Precambrian Research, 259, 243–261. [Google Scholar]
- 48. Tkachev, A. V. , & Rundqvist, D. V. (2016). Global trends in the evolution of metallogenic processes as a reflection of supercontinent cyclicity. Geology of Ore Deposits, 58, 263–283. [Google Scholar]
- 49. Yale, L. B. , & Carpenter, S. J. (1998). Large igneous provinces and giant dike swarms: Proxies for supercontinent cyclicity and mantle convection. Earth and Planetary Science Letters, 163, 109–122. [Google Scholar]
- 50. Li, Z.‐X. , & Zhong, S. (2009). Supercontinent–superplume coupling, true polar wander and plume mobility: Plate dominance in whole‐mantle tectonics. Physics of the Earth and Planetary Interiors, 176, 143–156. [Google Scholar]
- 51. Ernst, R. E. , Bleeker, W. , Söderlund, U. , & Kerr, A. C. (2013). Large igneous provinces and supercontinents: Toward completing the plate tectonic revolution. Lithos, 174, 1–14. [Google Scholar]
- 52. Condie, K. , Pisarevsky, S. A. , Korenaga, J. , & Gardoll, S. (2015). Is the rate of supercontinent assembly changing with time? Precambrian Research, 259, 278–289. [Google Scholar]
- 53. Söderlund, U. , Klausen, M. B. , Ernst, R. E. , & Bleeker, W. (2016). New advances in using large igneous provinces (LIPs) to reconstruct ancient supercontinents. Geologiska Foereningens I Stockholm Foerhandlingar, 138, 1–5. [Google Scholar]
- 54. Zhong, S. , Zhang, N. , Li, Z.‐X. , & Roberts, J. H. (2007). Supercontinent cycles, true polar wander, and very long‐wavelength mantle convection. Earth and Planetary Science Letters, 261, 551–564. [Google Scholar]
- 55. Yoshida, M. , & Santosh, M. (2011). Supercontinents, mantle dynamics and plate tectonics: A perspective based on conceptual vs. numerical models. Earth‐Science Reviews, 105, 1–24. [Google Scholar]
- 56. Ganne, J. , Feng, X. , Rey, P. , & De Andrade, V. (2016). Statistical petrology reveals a link between supercontinents cycle and mantle global climate. American Mineralogist, 101, 2768–2773. [Google Scholar]
- 57. Trim, S. J. , & Lowman, J. P. (2016). Interaction between the supercontinent cycle and the evolution of intrinsically dense provinces in the deep mantle. Journal of Geophysical Research, Solid Earth, 121, 8941–8969. [Google Scholar]
- 58. Gamal El Dien, H. , Doucet, L. S. , Li, Z.‐X. , Cox, G. , & Mitchell, R. (2019). Global geochemical fingerprinting of plume intensity suggests coupling with the supercontinent cycle. Nature Communication, 10, 5270. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 59. Heron, P. J. (2019). Mantle plumes and mantle dynamics in the Wilson cycle. In Wilson, R. W. , Houseman, G. A. , McCaffrey, K. J. W. , Doré, A. G., & Buiter, S. J. H. (Eds.), Fifty years of the Wilson cycle concept in plate tectonics (pp. 87–103.) Geological Society, London, Special Publications. [Google Scholar]
- 60. Doucet, L. S. , Li, Z.‐X. , Ernst, R. E. , Kirscher, U. , El Dien, H. G. , & Mitchell, R. N. (2020). Coupled supercontinent–mantle plume events evidenced by oceanic plume record. Geology, 48, 159–163. [Google Scholar]
- 61. Worsley, T. R. , Nance, D. , & Moody, J. B. (1984). Global tectonics and eustasy for the past 2 billion years. Marine Geology, 58, 373–400. [Google Scholar]
- 62. Heller, P. L. , & Angevine, C. L. (1985). Sea‐level cycles during the growth of Atlantic‐type oceans. Earth and Planetary Science Letters, 75, 417–426. [Google Scholar]
- 63. Cogné, J.‐P. , & Humler, E. (2008). Global scale patterns of continental fragmentation: Wilson's cycles as a constraint for long‐term sea‐level changes. Earth and Planetary Science Letters, 273, 251–259. [Google Scholar]
- 64. Conrad, C. P. (2013). The solid Earth's influence on sea level. Geological Society of America Bulletin, 125, 1027–1052. [Google Scholar]
- 65. Karlsen, K. S. , Domeier, M. , Gaina, C. , & Conrad, C. P. (2020). A tracer‐based algorithm for automatic generation of seafloor age grids fromplate tectonic reconstructions. Computers & Geosciences, 140, 104508. [Google Scholar]
- 66. Young, A. , Flament, N. , Williams, S. E. , Merdith, A. , Cao, X. , & Müller, R. D. (2022). Long‐term Phanerozoic sea level change from solid Earth processes. Earth and Planetary Science Letters, 584, 117451. [Google Scholar]
- 67. Horita, J. , Zimmermann, H. , & Holland, H. D. (2002). Chemical evolution of seawater during the Phanerozoic: Implications from the record of marine evaporates. Geochimica et Cosmochimica Acta, 66, 3733–3756. [Google Scholar]
- 68. Müller, R. D. , Dutkiewicz, A. , Seton, M. , & Gaina, C. (2013). Seawater chemistry driven by supercontinent assembly, breakup, and dispersal. Geology, 41, 907–910. [Google Scholar]
- 69. Goddéris, Y. , Donnadieu, Y. , Hir, G. L. , Lefebvre, V. , & Nardin, E. (2014). The role of palaeogeography in the Phanerozoic history of atmospheric CO2 and climate. Earth‐Science Reviews, 128, 122–138. [Google Scholar]
- 70. Condie, K. (2001). Precambrian superplumes and supercontinents: A record in black shales, carbon isotopes, and paleoclimates? Precambrian Research, 106, 239–260. [Google Scholar]
- 71. Shields, G. A. (2007). A normalised seawater strontium isotope curve: Possible implications for Neoproterozoic‐Cambrian weathering rates and the further oxygenation of the Earth. eEarth, 2, 35–42. [Google Scholar]
- 72. Algeo, T. J. , Luo, G. M. , Song, H. Y. , Lyons, T. W. , & Canfield, D. E. (2015). Reconstruction of secular variation in seawater sulfate concentrations. Biogeosciences, 12, 2131–2151. [Google Scholar]
- 73. Krapez, B. (1997). Sequence‐stratigraphic concepts applied to the identification of depositional basins and global tectonic cycles. Australian Journal of Earth Sciences, 44, 1–36. [Google Scholar]
- 74. Eriksson, P. G. , Catuneanu, O. , Nelson, D. R. , & Popa, M. (2005). Controls on Precambrian sea level change and sedimentary cyclicity. Sedimentary Geology, 176, 43–65. [Google Scholar]
- 75. Eriksson, P. G. , Banerjee, S. , Catuneanu, O. , Corcoran, P. L. , Eriksson, K. A. , Hiatt, E. E. , Laflamme, M. , Lenhardt, N. , Long, D. G. F. , Miall, A. D. , Mints, M. V. , Pufahl, P. K. , Sarkar, S. , Simpson, E. L. , & Williams, G. E. (2013). Secular changes in sedimentation systems and sequence stratigraphy. Gondwana Research, 24, 468–489. [Google Scholar]
- 76. Worsley, T. R. , & Kidder, D. L. (1991). First‐order coupling of paleogeography and CO2, with global surface temperature and its latitudinal contrast. Geology, 19, 1161–1164. [Google Scholar]
- 77. Lindsay, J. F. , & Brasier, M. D. (2004). The evolution of the Precambrian atmosphere: Carbon isotopic evidence from the Australian continent. In Eriksson, P. G. , Altermann, W. , Nelson, D. R. , Mueller, W. U., & Catuneanu, O. (Eds.), The Precambrian Earth: Tempos and events (pp. 388–403). Amsterdam: Elsevier. [Google Scholar]
- 78. Campbell, I. H. , & Allen, C. M. (2008). Formation of supercontinents linked to increases in atmospheric oxygen. Nature Geoscience, 1, 554–558. [Google Scholar]
- 79. Nance, R. D. , Worsley, T. R. , & Moody, J. B. (1986). Post‐Archean biogeochemical cycles and long‐term episodicity in tectonic process. Geology, 14, 514–518. [Google Scholar]
- 80. Santosh, M. (2010). Supercontinent tectonics and biogeochemical cycle: A matter of ‘life and death’. Geoscience Frontiers, 1, 21–30. [Google Scholar]
- 81. Hoffman, P. F. (1999). The break‐up of Rodinia, birth of Gondwana, true polar wander and the snowball Earth. Journal of African Earth Sciences, 28, 17–33. [Google Scholar]
- 82. Eyles, N. (2008). Glacio‐epochs and the supercontinent cycle after ∼ 3.0 Ga: Tectonic boundary conditions for glaciation. Palaeogeography, Palaeoclimatology, Palaeoecology, 258, 89–129. [Google Scholar]
- 83. Young, G. M. (2013). Precambrian supercontinents, glaciations, atmospheric oxygenation, metazoan evolution and an impact that may have changed the second half of Earth history. Geoscience Frontiers, 4, 247–261. [Google Scholar]
- 84. Jellinek, A. M. , Lenardic, A. , & Pierrehumbert, R. T. (2020). Ice, fire or fizzle: The climate footprint of Earth's supercontinental cycles. Geochemistry, Geophysics, Geosystems, 21, e2019GC008464. [Google Scholar]
- 85. Valentine, J. W. , & Moores, E. M. (1970). Plate tectonic regulation of faunal diversity and sea level. Nature, 228, 657–659. [DOI] [PubMed] [Google Scholar]
- 86. Hannisdal, B. , & Peters, S. E. (2011). Phanerozoic Earth system evolution and marine biodiversity. Science, 334, 1121–1124. [DOI] [PubMed] [Google Scholar]
- 87. Lindsay, J. F. , & Brasier, M. D. (2002). Did global tectonics drive early biosphere evolution? Carbon isotope record from 2.6 to 1.9 Ga carbonates of Western Australian basins. Precambrian Research, 114, 1–34. [Google Scholar]
- 88. Bond, D. P. G. , & Grasby, S. E. (2017). On the causes of mass extinctions. Palaeogeography, Palaeoclimatology, Palaeoecology, 478, 3–29. [Google Scholar]
- 89. Condie, K. C. (2011). The supercontinent cycle. Chapter 8 in Earth as an evolving planetary system (2nd ed., pp. 317–355). Amsterdam: Academic Press. [Google Scholar]
- 90. Bradley, D. C. (2011). Secular trends in the geologic record and the supercontinent cycle. Earth‐Science Reviews, 108, 16–33. [Google Scholar]
- 91. Nance, R. D. , & Murphy, J. B. (2013). Origins of the supercontinent cycle. Geoscience Frontiers, 4, 439–448. [Google Scholar]
- 92. Nance, R. D. , Worsley, T. R. , & Moody, J. B. (1988). The supercontinent cycle. Scientific American, 259, 72–79.3072674 [Google Scholar]
- 93. Umbgrove, J. H. F. (1947). The pulse of the Earth. The Hague, Netherlands: Martinus Nijhoff. [Google Scholar]
- 94. Holmes, A. (1951). The sequence of Precambrian orogenic belts in south and central Africa. In International Geological Congress. London. [Google Scholar]
- 95. Holmes, A. (1954). Principles of physical geology. London: Thomas Nelson and Sons. [Google Scholar]
- 96. Wilson, A. F. , Compston, W. , Jeffery, P. M. , & Riley, G. H. (1959). Radiometric ages from the Precambrian rocks in Australia. Journal of the Geological Society of Australia, 6, 179–195. [Google Scholar]
- 97. Gastil, R. G. (1960). The distribution of mineral dates in time and space. American Journal of Science, 258, 1–35. [Google Scholar]
- 98. Runcorn, S. K. (1962). Convection currents in the Earth's mantle. Nature, 195, 1248–1249. [Google Scholar]
- 99. Sloss, L. L. (1963). Sequences in the cratonic interior of North America. Geological Society of America Bulletin, 74, 93–114. [Google Scholar]
- 100. Sutton, J. (1963). Long‐term cycles in the evolution of the continents. Nature, 198, 731–735. [Google Scholar]
- 101. Santosh, M. , & Zhao, G. (2009). Supercontinent dynamics. Gondwana Research, 15, 225–227. [Google Scholar]
- 102. Rogers, J. J. W. , & Santosh, M. (2004). Continents and supercontinents. New York: Oxford University Press. [Google Scholar]
- 103. Goldfarb, R. J. , Bradley, D. , & Leach, D. L. (2010). Secular variation in economic geology. Economic Geology, 105, 459–465. [Google Scholar]
- 104. Santosh, M. (2010). A synopsis of recent conceptual models on supercontinent tectonics in relation to mantle dynamics, life evolution and surface environment. Journal of Geodynamics, 50, 116–133. [Google Scholar]
- 105. Strand, K. (2012). Global and continental‐scale glaciations on the Precambrian earth. Marine and Petroleum Geology, 33, 69–79. [Google Scholar]
- 106. Cawood, P. A. , Hawkesworth, C. J. , & Dhuime, B. (2013). The continental record and the generation of continental crust. Geological Society of America Bulletin, 125, 14–32. [Google Scholar]
- 107. Spencer, C. J. , Cawood, P. A. , Hawkesworth, C. J. , Raub, T. D. , Prave, A. R. , & Roberts, N. M. W. (2014). Proterozoic onset of crustal reworking and collisional tectonics: Reappraisal of the zircon oxygen isotope record. Geology, 42, 451–454. [Google Scholar]
- 108. Cao, W. , Lee, C.‐T. y. A. , & Lackey, J. S. (2017). Episodic nature of continental arc activity since 750 Ma: A global compilation. Earth and Planetary Science Letters, 461, 85–95. [Google Scholar]
- 109. Brown, M. (2014). The contribution of metamorphic petrology to understanding lithosphere evolution and geodynamics. Geoscience Frontiers, 5, 553–569. [Google Scholar]
- 110. Anderson, D. L. (1982). Hotspots, polar wander, Mesozoic convection and the geoid. Nature, 297, 391–393. [Google Scholar]
- 111. Coltice, N. , Bertrand, H. , Rey, P. , Jourdan, F. , Phillips, B. R. , & Ricard, Y. (2009). Global warming of the mantle beneath continents back to the Archaean. Gondwana Research, 15, 254–266. [Google Scholar]
- 112. Miller, K. G. , Kominz, M. A. , Browning, J. V. , Wright, J. D. , Mountain, G. S. , Katz, M. E. , Sugarman, P. J. , Cramer, B. S. , Christie‐Blick, N. , & Pekar, S. F. (2005). The Phanerozoic record of global sea‐level change. Science, 310, 1293–1298. [DOI] [PubMed] [Google Scholar]
- 113. Guillaume, B. , Pochat, S. , Monteux, J. , Husson, L. , & Choblet, G. (2016). Can eustatic charts go beyond first order? Insights from the Permian–Triassic. Lithosphere, 8, 505–518. [Google Scholar]
- 114. Kump, L. R. , Brantley, S. L. , & Arthur, M. A. (2000). Chemical weathering, atmospheric CO2, and climate. Annual Review of Earth and Planetary Sciences, 28, 611–667. [Google Scholar]
- 115. Scotese, C. R. , Song, H. , Mills, B. J. W. , & Van Der Meer, D. G. (2021). Phanerozoic paleotemperatures: The earth's changing climate during the last 540 million years. Earth‐Science Reviews, 215, 103503. [Google Scholar]
- 116. Goddéris, Y. , Le Hir, G. , Macouin, M. , Donnadieu, Y. , Hubert‐Théou, L. , Dera, G. , Aretz, M. , Fluteau, F. , Li, Z. X. , & Halverson, G. P. (2017). Paleogeographic forcing of the strontium isotopic cycle in the Neoproterozoic. Gondwana Research, 42, 151–162. [Google Scholar]
- 117. Paulsen, T. , Deering, C. , Sliwinski, J. , Chatterjee, S. , & Bachman, O. (2022). Continental magmatism and uplift as the primary driver for first‐order oceanic 87Sr/86Sr variability with implications for global climate and atmospheric oxygenation. GSA Today, 32, 4–10. [Google Scholar]
- 118. Bercovici, D. , & Long, M. D. (2014). Slab rollback instability and supercontinent dispersal. Geophysical Research Letters, 41, 6659–6666. [Google Scholar]
- 119. Keppie, F. (2015). How subduction broke up Pangaea with implications for the supercontinent cycle. In Li, Z.‐X. , Evans, D. A. D. , & Murphy, J. B. (Eds.), Supercontinent cycles through Earth history (pp. 265–288). Geological Society, London, Special Publications. [Google Scholar]
- 120. Dal Zilio, L. , Faccenda, M. , & Capitanio, F. (2018). The role of deep subduction in supercontinent breakup. Tectonophysics, 746, 312–324. [Google Scholar]
- 121. Ernst, R. E. , Wingate, M. T. D. , Buchan, K. L. , & Li, Z. X. (2008). Global record of 1600–700 Ma large igneous provinces (LIPs): Implications for the reconstruction of the proposed Nuna (Columbia) and Rodinia supercontinents. Precambrian Research, 160, 159–178. [Google Scholar]
- 122. Bradley, D. C. (2008). Passive margins through Earth history. Earth‐Science Reviews, 91, 1–26. [Google Scholar]
- 123. Kirschner, J. P. , Kominz, M. A. , & Mwakanyamale, K. E. (2010). Quantifying extension of passive margins: Implications for sea level change. Tectonics, 29, TC4005. [Google Scholar]
- 124. Craig, J. , Thurow, J. , Thusu, B. , Whitham, A. , & Abutarruma, Y. (2009). Global Neoproterozoic petroleum systems: The emerging potential in North Africa. In Craig, J., Thurow, J., Thusu, B., Whitham, A., & Abutarrumam Y. (Eds.), Global Neoproterozoic petroleum systems: The emerging potential in North Africa (pp. 1–25). Geological Society, London, Special Publications. [Google Scholar]
- 125. Gurnis, M. (1988). Large‐scale mantle convection and the aggregation and dispersal of supercontinents. Nature, 332, 695–699. [Google Scholar]
- 126. Lowman, J. P. , & Jarvis, G. T. (1999). Effects of mantle heat source distribution on supercontinent stability. Journal of Geophysical Research, Solid Earth, 104, 12733–12746. [Google Scholar]
- 127. Murphy, J. B. , & Nance, R. D. (2003). Do supercontinents introvert or extrovert?: Sm‐Nd isotope evidence. Geology, 31, 873–876. [Google Scholar]
- 128. Vaughan, A. P. M. , & Storey, B. C. (2007). A new supercontinent self‐destruct mechanism: Evidence from the Late Triassic–Early Jurassic. Journal of the Geological Society of London, 164, 383–392. [Google Scholar]
- 129. Padma Rao, B. , & Ravi Kumar, M. (2014). Seismic evidence for slab graveyards atop the core mantle boundary beneath the Indian Ocean Geoid Low. Physics of the Earth and Planetary Interiors, 236, 52–59. [Google Scholar]
- 130. Voosen, P. (2016). Graveyard of cold slabs mapped in Earth's mantle. Science, 354, 954–955. [DOI] [PubMed] [Google Scholar]
- 131. Heron, P. J. , Lowman, J. P. , & Stein, C. (2015). Influences on the positioning of mantle plumes following supercontinent formation. Journal of Geophysical Research, Solid Earth, 120, 3628–3648. [Google Scholar]
- 132. Evans, D. A. D. (2003). True polar wander and supercontinents. Tectonophysics, 362, 303–320. [Google Scholar]
- 133. Mitchell, R. N. , Kilian, T. M. , & Evans, D. A. D. (2012). Supercontinent cycles and the calculation of absolute palaeolongitude in deep time. Nature, 482, 208–211. [DOI] [PubMed] [Google Scholar]
- 134. Collins, W. J. (2003). Slab pull, mantle convection, and Pangaean assembly and dispersal. Earth and Planetary Science Letters, 205, 225–237. [Google Scholar]
- 135. Zhang, N. , Dang, Z. , Huang, C. , & Li, Z.‐X. (2018). The dominant driving force for supercontinent breakup: Plume push or subduction retreat? Geoscience Frontiers, 9, 997–1007. [Google Scholar]
- 136. Niu, Y. (2020). On the cause of continental breakup: A simple analysis in terms of driving mechanisms of plate tectonics and mantle plumes. Journal of Asian Earth Sciences, 194, 104367. [Google Scholar]
- 137. Condie, K. C. , Davaille, A. , Aster, R. C. , & Arndt, N. (2014). Upstairs‐downstairs: Supercontinents and large igneous provinces, are they related? International Geology Review, 57, 1341–1348. [Google Scholar]
- 138. Condie, K. C. , & Puetz, S. J. (2019). Time series analysis of mantle cycles Part II: The geologic record in zircons, large igneous provinces and mantle lithosphere. Geoscience Frontiers, 10, 1327–1336. [Google Scholar]
- 139. Raymo, M. E. , & Ruddiman, W. F. (1992). Tectonic forcing of late Cenozoic climate. Nature, 359, 117–122. [Google Scholar]
- 140. Goddéris, Y. , Donnadieu, Y. , Lefebvre, V. , Le Hir, G. , & Nardin, E. (2012). Tectonic control of continental weathering, atmospheric CO2, and climate over Phanerozoic times. Comptes Rendus Geosciences, 344, 652–662. [Google Scholar]
- 141. Chamberlin, T. C. (1899). An attempt to frame a working hypothesis of the cause of glacial periods on an atmospheric basis. Journal of Geology, 7, 545–584. [Google Scholar]
- 142. Ernst, R. E. , Bond, D. P. G. , Zhang, S.‐H. , Buchan, K. L. , Grasby, S. E. , Youbi, N. , El Bilali, H. , Bekker, A. , & Doucet, L. S. (2021). Large igneous province record through time and implications for secular environmental changes and geological time‐scale boundaries. In Ernst, R. E. , Dickson, A. J. , & Bekker, A. (Eds.), Large igneous provinces: A driver of global environmental and biotic changes (pp. 1–26). AGU Geophysical Monograph. [Google Scholar]
- 143. Ernst, R. E. (2014). Large igneous provinces. Cambridge University Press. [Google Scholar]
- 144. Ernst, R. E. , & Youbi, N. (2017). How large igneous provinces affect global climate, sometimes cause mass extinctions, and represent natural markers in the geological record. Palaeogeography, Palaeoclimatology, Palaeoecology, 478, 30–52. [Google Scholar]
- 145. Donnadieu, Y. , Goddéris, Y. , Ramstein, G. , Nédélec, A. , & Meert, J. (2004). A ‘snowball Earth’ climate triggered by continental break‐up through changes in runoff. Nature, 428, 303–306. [DOI] [PubMed] [Google Scholar]
- 146. Vail, P. R. , Mitchum, R. M. Jr. , & Thompson, S. III . (1977). Seismic stratigraphy and global changes of sea level, Part 4: Global cycles of relative changes of sea level. In Payton, C. E. (Ed.), Seismic stratigraphy – Applications to hydrocarbon exploration (pp. 83–97). American Association of Petroleum Geologists. [Google Scholar]
- 147. Hallam, A. (1992). Phanerozoic sea level changes. New York: Columbia University Press. [Google Scholar]
- 148. Berner, R. A. (1999). A new look at the long‐term carbon cycle. GSA Today, 9, 2–6. [Google Scholar]
- 149. Berner, R. A. (2004). The Phanerozoic carbon cycle: CO2 and O2 . Oxford: Oxford University Press. [Google Scholar]
- 150. Berner, R. A. (2001). Geocarb III: A revised model of atmospheric CO2 over Phanerozoic time. American Journal of Science, 301, 182–204. [DOI] [PubMed] [Google Scholar]
- 151. Gaillardet, J. , Dupré, B. , Louvat, P. , & Allègre, C. J. (1999). Global silicate weathering and CO2 consumption rates deduced from the chemistry of large rivers. Chemical Geology, 159, 3–30. [Google Scholar]
- 152. Brady, P. V. , & Gíslason, S. R. (1997). Seafloor weathering controls on atmospheric CO2 and global climate. Geochimica et Cosmochimica Acta, 61, 965–973. [Google Scholar]
- 153. Gillis, K. M. , & Coogan, L. A. (2011). Secular variation in carbon uptake into the ocean crust. Earth and Planetary Science Letters, 302, 385–392. [Google Scholar]
- 154. Coogan, L. A. , & Dosso, S. E. (2015). Alteration of ocean crust provides a strong temperature dependent feedback on the geological carbon cycle and is a primary driver of the Sr‐isotopic composition of seawater. Earth and Planetary Science Letters, 415, 38–46. [Google Scholar]
- 155. Lee, C.‐T. A. , Shen, B. , Slotnick, B. S. , Liao, K. , Dickens, G. R. , Yokoyama, Y. , Lenardic, A. , Dasgupta, R. , Jellinek, M. , Lackey, J. S. , Schneider, T. , & Tice, M. M. (2013). Continental arc–island arc fluctuations, growth of crustal carbonates, and long‐term climate change. Geosphere, 9, 21–36. [Google Scholar]
- 156. West, A. J. (2012). Thickness of the chemical weathering zone and implications for erosional and climatic drivers of weathering and for carbon‐cycle feedbacks. Geology, 40, 811–814. [Google Scholar]
- 157. Fielding, C. R. , Frank, T. D. , & Isbell, J. L. (2008). The late Palaeozoic ice age: A review of current understanding and synthesis of global climate patterns. In Fielding, C. R. , Frank, T. D. , & Isbell, J. L. (Eds.), Resolving the late Paleozoic ice age in time and space (pp. 343–354). Geological Society of America. [Google Scholar]
- 158. Montañez, I. P. , & Poulsen, C. J. (2013). The Late Palaeozoic Ice Age: An evolving paradigm. Annual Review of Earth and Planetary Sciences, 41, 629–656. [Google Scholar]
- 159. Prave, A. R. , Condon, D. J. , Hoffmann, K. H. , Tapster, S. , & Fallick, A. E. (2016). Duration and nature of the end‐Cryogenian (Marinoan) glaciation. Geology, 44, 631–634. [Google Scholar]
- 160. Hoffman, P. F. , & Li, Z.‐X. (2009). A palaeogeographic context for Neoproterozoic glaciations. Palaeogeography, Palaeoclimatology, Palaeoecology, 277, 158–172. [Google Scholar]
- 161. Hebert, C. L. , Kaufman, A. J. , Penniston‐Dorland, S. C. , & Martin, A. J. (2010). Radiometric and stratigraphic constraints on terminal Ediacaran (post‐Gaskiers) glaciation and metazoan evolution. Precambrian Research, 182, 402–412. [Google Scholar]
- 162. Pu, J. P. , Bowring, S. A. , Ramezani, J. , Myrow, P. , Raub, T. D. , Landing, E. d. , Mills, A. , Hodgin, E. , & Macdonald, F. A. (2016). Dodging snowballs: Geochronology of the Gaskiers glaciation and the first appearance of the Ediacaran biota. Geology, 44, 955–958. [Google Scholar]
- 163. Hebert, C. L. , Kaufman, A. J. , Penniston‐Dorland, S. C. , & Martin, A. J. (2018). A ∼565 Ma old glaciation in the Ediacaran of peri‐Gondwanan West Africa. International Journal of Earth Sciences, 182, 402–412. [Google Scholar]
- 164. Rooney, A. D. , Strauss, J. V. , Brandon, A. D. , & Macdonald, F. A. (2015). A Cryogenian chronology: Two long‐lasting, synchronous Neoproterozoic snowball Earth glaciations. Geology, 43, 459–462. [Google Scholar]
- 165. Cox, G. M. , Isakson, V. , Hoffman, P. F. , Gernon, T. M. , Schmitz, M. D. , Shahin, S. , Collins, A. S. , Preiss, W. , Blades, M. L. , Mitchell, R. N. , & Nordsvan, A. (2018). South Australian U‐Pb zircon (CA‐ID‐TIMS) age supports globally synchronous Sturtian deglaciation. Precambrian Research, 315, 257–263. [Google Scholar]
- 166. Lan, Z. , Huyskens, M. H. , Lu, K. , Li, X.‐H. , Zhang, G. , Lu, D. , & Yin, Q.‐Z. (2020). Toward refining the onset age of Sturtian glaciation in South China. Precambrian Research, 338, 105555. [Google Scholar]
- 167. Brasier, A. T. , Martin, A. P. , Melezhik, V. A. , Prave, A. R. , Condon, D. J. , & Fallick, A. E. (2013). Earth's earliest global glaciation? Carbonate geochemistry and geochronology of the Polisarka Sedimentary Formation, Kola Peninsula, Russia. Precambrian Research, 235, 278–294. [Google Scholar]
- 168. Rasmussen, B. , Bekker, A. , & Fletcher, I. R. (2013). Correlation of Paleoproterozoic glaciations based on U‐Pb zircon ages for tuff beds in the Transvaal and Huronian Supergroups. Earth and Planetary Science Letters, 382, 173–180. [Google Scholar]
- 169. Tang, H. , & Chen, Y. (2013). Global glaciations and atmospheric change at ca. 2.3 Ga. Geoscience Frontiers, 4, 583–596. [Google Scholar]
- 170. Von Brunn, V. , & Gold, D. J. C. (1993). Diamictite in the Archaean Pongola sequence of southern Africa. Journal of African Earth Sciences, 16, 367–374. [Google Scholar]
- 171. Young, G. M. , Brunn, V. V. , Gold, D. J. C. , & Minter, W. E. L. (1998). Earth's oldest reported glaciation: Physical and chemical evidence from the Archean Mozaan Group (∼2.9 Ga) of South Africa. Journal of Geology, 106, 523–538. [Google Scholar]
- 172. Luskin, C. , Wilson, A. , Gold, D. , & Hofmann, A. (2019). The Pongola Supergroup: Mesoarchaean deposition following Kaapvaal Craton stabilization. In Kröner, A. , & Hofmann, A. (Eds.), The Archaean geology of the Kaapvaal Craton, Southern Africa (pp. 225–254). Amsterdam: Springer. [Google Scholar]
- 173. Young, G. M. (2019). Aspects of the Archean‐Proterozoic transition: How the great Huronian Glacial Event was initiated by rift‐related uplift and terminated at the rift‐drift transition during break‐up of Lauroscandia. Earth‐Science Reviews, 190, 171–189. [Google Scholar]
- 174. Eyles, N. (1993). Earth's glacial record and its tectonic setting. Earth‐Science Reviews, 35, 1–248. [Google Scholar]
- 175. Young, G. M. (1995). Are Neoproterozoic glacial deposits preserved on the margins of Laurentia related to the fragmentation of two supercontinents? Geology, 23, 153–156. [Google Scholar]
- 176. Delabroye, A. , & Vecoli, M. (2010). The end‐Ordovician glaciation and the Hirnantian Stage: A global review and questions about Late Ordovician event stratigraphy. Earth‐Science Reviews, 98, 269–282. [Google Scholar]
- 177. Finlay, A. J. , Selby, D. , & Gröcke, D. R. (2010). Tracking the Hirnantian glaciation using Os isotopes. Earth and Planetary Science Letters, 293, 339–348. [Google Scholar]
- 178. Gumsley, A. P. , Chamberlain, K. R. , Bleeker, W. , Söderlund, U. , De Kock, M. O. , Larsson, E. R. , & Bekker, A. (2017). Timing and tempo of the Great Oxidation Event. Proceedings of the National Academy of Sciences of the United States of America, 114, 1811–1816. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 179. Holland, H. D. (2002). Volcanic gases, black smokers, and the Great Oxidation Event. Geochimica et Cosmochimica Acta, 66, 3811–3826. [Google Scholar]
- 180. Schopf, J. W. (2014). Geological evidence of oxygenic photosynthesis and the biotic response to the 2400–2200 Ma “Great Oxidation Event.” Biochemistry Moscow, 79, 165–177. [DOI] [PubMed] [Google Scholar]
- 181. Ciborowski, T. J. R. , & Kerr, A. C. (2016). Did mantle plume magmatism help trigger the Great Oxidation Event? Lithos, 246–247, 128–133. [Google Scholar]
- 182. Beukes, N. J. , Dorland, H. , Gutzmer, J. , Nedachi, M. , & Ohmoto, H. (2002). Tropical laterites, life on land, and the history of atmospheric oxygen in the Paleoproterozoic. Geology, 30, 491–494. [Google Scholar]
- 183. Gaucher, C. , & Frei, R. (2018). The Archean‐Proterozoic boundary and the Great Oxidation Event. In Sial, A. N. , Gaucher, C. , Ramkumar, M. , & Ferreira, V. P. (Eds.), Chemostratigraphy across major chronological boundaries (pp. 35–45). American Geophysical Union. [Google Scholar]
- 184. Kump, L. R. , Fallick, A. E. , Melezhik, V. A. , Strauss, H. , & Lepland, A. (2013). The Great Oxidation Event. In Melezhik, V. , Prave, A. R. , Hanski, E. J. , Fallick, A. E., Lepland, A., Kump, L. R., & Strauss, H. (Eds.), Reading the archive of Earth's oxygenation. Frontiers in Earth sciences (pp. 1517–1533). Heidelberg: Springer. [Google Scholar]
- 185. Eriksson, P. G. , & Cheney, E. S. (1992). Evidence for the transition to an oxygen‐rich atmosphere during the evolution of red beds in the Lower Proterozoic sequences of southern Africa. Precambrian Research, 54, 257–269. [Google Scholar]
- 186. Farquhar, J. , Bao, H. , & Thiemens, M. (2000). Atmospheric influence of Earth's earliest sulfur cycle. Science, 289, 756–758. [DOI] [PubMed] [Google Scholar]
- 187. Hoffman, P. F. (2013). The Great Oxidation and a Siderian snowball Earth: MIF‐S based correlation of Paleoproterozoic glacial epochs. Chemical Geology, 362, 143–156. [Google Scholar]
- 188. Bekker, A. , Kaufman, A. , Karhu, J. , & Eriksson, K. (2005). Evidence for Paleoproterozoic cap carbonates in North America. Precambrian Research, 137, 167–206. [Google Scholar]
- 189. Kopp, R. E. , Kirschvink, J. L. , Hilburn, I. A. , & Nash, C. Z. (2005). The Paleoproterozoic snowball Earth: A climate disaster triggered by the evolution of oxygenic photosynthesis. Proceedings of the National Academy of Sciences of the United States of America, 102, 11131–11136. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 190. Tajika, E. , & Harada, M. (2019). Great Oxidation Event and snowball Earth. In Yamagishi, A. , Kakegawa, T. , & Usui, T. (Eds.), Astrobiology (pp. 261–271). Singapore: Springer. [Google Scholar]
- 191. Kirschvink, J. L. (1992). Late Proterozoic low‐latitude global glaciation: The snowball Earth. In Schopf, J. W. , & Klein, C. (Eds.), The Proterozoic biosphere: A multidisciplinary study (pp. 51–52). Cambridge University Press. [Google Scholar]
- 192. Hoffman, P. F. , & Schrag, D. P. (2002). The snowball Earth hypothesis: Testing the limits of global change. Terra Nova, 14, 129–155. [Google Scholar]
- 193. Budyko, M. I. (1969). The effect of solar radiation variations on the climate of the Earth. Tellus, 21, 611–619. [Google Scholar]
- 194. Sellers, W. D. (1969). A global climatic model based on the energy balance of the Earth‐atmosphere system. Journal of Applied Meteorology, 8, 392–400. [Google Scholar]
- 195. Hoffman, P. F. , Abbot, D. S. , Ashkenazy, Y. , Benn, D. I. , Brocks, J. J. , Cohen, P. A. , Cox, G. M. , Creveling, J. R. , Donnadieu, Y. , Erwin, D. H. , Fairchild, I. J. , Ferreira, D. , Goodman, J. C. , Halverson, G. P. , Jansen, M. F. , Le Hir, G. , Love, G. D. , Macdonald, F. A. , Maloof, A. C. , … Warren, S. G. (2017). Snowball Earth climate dynamics and Cryogenian geology‐geobiology. Science Advances, 3, e1600983. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 196. Macdonald, F. A. , Schmitz, M. D. , Strauss, J. V. , Halverson, G. P. , Gibson, T. M. , Eyster, A. , Cox, G. , Mamrol, P. , & Crowley, J. L. (2018). Cryogenian of Yukon. Precambrian Research, 319, 114–143. [Google Scholar]
- 197. Caldeira, K. , & Kasting, J. F. (1992). Susceptibility of the early Earth to irreversible glaciation caused by carbon dioxide clouds. Nature, 359, 226–228. [DOI] [PubMed] [Google Scholar]
- 198. Kirschvink, J. L. , & Raub, T. D. (2003). A methane fuse for the Cambrian explosion: Carbon cycles and true polar wander. Comptes Rendus Geoscience, 335, 65–78. [Google Scholar]
- 199. Condie, K. C. (2004). Supercontinents and superplume events: Distinguishing signals in the geologic record. Physics of the Earth and Planetary Interiors, 146, 319–332. [Google Scholar]
- 200. Coppold, M. , & Powell, W. (2006). A geoscience guide to the Burgess Shale: Geology and paleontology in Yoho National Park (2nd ed.). Field, BC: Burgess Shale Geoscience Foundation. [Google Scholar]
- 201. Hay, W. W. (2016). Experimenting on a small planet: A history of scientific discoveries, a future of climate change and global warming (2nd ed.). Basel: Springer. [Google Scholar]
- 202. Cohen, A. S. , Coe, A. L. , Harding, S. M. , & Schwark, L. (2004). Osmium isotope evidence for the regulation of atmospheric CO2 by continental weathering. Geology, 32, 157–160. [Google Scholar]
- 203. Wignall, P. B. (2005). The timing of paleoenvironmental change and cause‐and‐effect relationships during the early Jurassic mass extinction in Europe. American Journal of Science, 305, 1014–1032. [Google Scholar]
- 204. Poulsen, C. J. , Tabor, C. , & White, J. D. (2015). Long‐term climate forcing by atmospheric oxygen concentrations. Science, 348, 1238–1241. [DOI] [PubMed] [Google Scholar]
- 205. Bryan, S. E. , & Ferrari, L. (2013). Large igneous provinces and silicic large igneous provinces: Progress in our understanding over the last 25 years. Geological Society of America Bulletin, 125, 1053–1078. [Google Scholar]
- 206. Wang, Y. U. , Santosh, M. , Luo, Z. , & Hao, J. (2015). Large igneous provinces linked to supercontinent assembly. Journal of Geodynamics, 85, 1–10. [Google Scholar]
- 207. Dalziel, I. W. D. , Lawver, L. A. , & Murphy, J. B. (2000). Plumes, orogenesis, and supercontinental fragmentation. Earth and Planetary Science Letters, 178, 1–11. [Google Scholar]
- 208. Santosh, M. , Maruyama, S. , & Yamamoto, S. (2009). The making and breaking of supercontinents: Some speculations based on superplumes, super downwelling and the role of tectosphere. Gondwana Research, 15, 324–341. [Google Scholar]
- 209. Ernst, R. , & Bleeker, W. (2010). Large igneous provinces (LIPs), giant dyke swarms, and mantle plumes: Significance for breakup events within Canada and adjacent regions from 2.5 Ga to the Present. Canadian Journal of Earth Sciences, 47, 695–739. [Google Scholar]
- 210. Klausen, M. B. (2020). Conditioned duality between supercontinental ‘assembly’ and ‘breakup’ LIPs. Geoscience Frontiers, 11, 1635–1649. [Google Scholar]
- 211. Pastor‐Galán, D. , Nance, R. D. , Murphy, J. B. , & Spencer, C. J. (2019). Supercontinents: Myths, mysteries, and milestones. In Wilson, R. W. , Houseman, G. A. , & McCaffrey, K. J. W. , Doré, A. G., & Buiter, S. J. H. (Eds.), Fifty years of the Wilson cycle concept in plate tectonics (pp. 39–64). Geological Society, London, Special Publications. [Google Scholar]
- 212. Puetz, S. J. , & Condie, K. C. (2019). Time series analysis of mantle cycles Part I: Periodicities and correlations among seven global isotopic databases. Geoscience Frontiers, 10, 1305–1326. [Google Scholar]
- 213. Condie, K. C. , Pisarevsky, S. A. , & Puetz, S. J. (2021). LIPs, orogens and supercontinents: The ongoing saga. Gondwana Research, 96, 105–121. [Google Scholar]
- 214. Torsvik, T. H. , Smethurst, M. A. , Burke, K. , & Steinberger, B. (2006). Large igneous provinces generated from the margins of the large low velocity provinces in the deep mantle. Geophysical Journal International, 167, 1447–1460. [Google Scholar]
- 215. Torsvik, T. H. , Burke, K. , Steinberger, B. , Webb, S. J. , & Ashwal, L. D. (2010). Diamonds sampled by plumes from the core–mantle boundary. Nature, 466, 352–355. [DOI] [PubMed] [Google Scholar]
- 216. Burke, K. , Steinberger, B. , Torsvik, T. H. , & Smethurst, M. A. (2008). Plume generation zones at the margins of large low shear velocity provinces on the core–mantle boundary. Earth and Planetary Science Letters, 265, 49–60. [Google Scholar]
- 217. Condie, K. C. (2001). Mantle plumes and their record in Earth history. Cambridge University Press. [Google Scholar]
- 218. Ernst, R. E. , & Buchan, K. L. (2003). Recognizing mantle plumes in the geological record. Annual Review of Earth and Planetary Sciences, 31, 469–523. [Google Scholar]
- 219. Guzewich, S. D. , Oman, L. D. , Richardson, J. A. , Whelley, P. L. , Bastelberger, S. T. , Young, K. E. , Bleacher, J. E. , Fauchez, T. J. , & Kopparapu, R. K. (2022). Volcanic climate warming through radiative and dynamical feedbacks of SO2 emissions. Geophysical Research Letters, 49, e2021GL096612. [Google Scholar]
- 220. Ganino, C. , & Arndt, N. T. (2009). Climate changes caused by degassing of sediments during the emplacement of large igneous provinces. Geology, 37, 323–326. [Google Scholar]
- 221. Wignall, P. B. (2001). Large igneous provinces and mass extinctions. Earth‐Science Reviews, 53, 1–33. [Google Scholar]
- 222. Courtillot, V. , & Renne, P. R. (2003). On the ages of flood basalt events. Comptes Rendus Geoscience, 335, 113–140. [Google Scholar]
- 223. Kiselev, A. I. , Ernst, R. E. , Yarmolyuk, V. V. , & Egorov, K. N. (2012). Radiated rifts and dyke swarms of the Middle Paleozoic Yakutsk plume of eastern Siberian craton. Journal of Asian Earth Sciences, 45, 1–16. [Google Scholar]
- 224. Ernst, R. E. , Rodygin, S. A. , & Grinev, O. M. (2020). Age correlation of large igneous provinces with Devonian biotic crises. Global and Planetary Change, 185, 103097. [Google Scholar]
- 225. Zhou, M.‐F. U. , Malpas, J. , Song, X.‐Y. , Robinson, P. T. , Sun, M. , Kennedy, A. K. , Lesher, C. M. , & Keays, R. R. (2002). A temporal link between the Emeishan large igneous province (SW China) and the end‐Guadalupian mass extinction. Earth and Planetary Science Letters, 196, 113–122. [Google Scholar]
- 226. Ivanov, A. V. , He, H. , Yan, L. , Ryabov, V. V. , Shevko, A. Y. , Palesskii, S. V. , & Nikolaeva, I. V. (2013). Siberian Traps large igneous province: Evidence for two flood basalt pulses around the Permo‐Triassic boundary and in the Middle Triassic, and contemporaneous granitic magmatism. Earth‐Science Reviews, 122, 58–76. [Google Scholar]
- 227. Black, B. A. , Neely, R. R. , Lamarque, J.‐F. , Elkins‐Tanton, L. T. , Kiehl, J. T. , Shields, C. A. , Mills, M. J. , & Bardeen, C. (2018). Systemic swings in end‐Permian climate from Siberian Traps carbon and sulfur outgassing. Nature Geoscience, 11, 949–954. [Google Scholar]
- 228. Blackburn, T. J. , Olsen, P. E. , Bowring, S. A. , Mclean, N. M. , Kent, D. V. , Puffer, J. , Mchone, G. , Rasbury, E. T. , & Et‐Touhami, M. (2013). Zircon U‐Pb geochronology links the end‐triassic extinction with the Central Atlantic Magmatic Province. Science, 340, 941–945. [DOI] [PubMed] [Google Scholar]
- 229. Percival, L. M. E. , Witt, M. L. I. , Mather, T. A. , Hermoso, M. , Jenkyns, H. C. , Hesselbo, S. P. , Al‐Suwaidi, A. H. , Storm, M. S. , Xu, W. , & Ruhl, M. (2015). Globally enhanced mercury deposition during the end‐Pliensbachian extinction and Toarcian OAE: A link to the Karoo–Ferrar Large Igneous Province. Earth and Planetary Science Letters, 428, 267–280. [Google Scholar]
- 230. Schoene, B. , Samperton, K. M. , Eddy, M. P. , Keller, G. , Adatte, T. , Bowring, S. A. , Khadri, S. F. R. , & Gertsch, B. (2015). U‐Pb geochronology of the Deccan Traps and relation to the end‐Cretaceous mass extinction. Science, 347, 182–184. [DOI] [PubMed] [Google Scholar]
- 231. Bond, D. P. G. , & Wignall, P. B. (2014). Large igneous provinces and mass extinctions: An update. In Keller, G. , & Kerr, A. C. (Eds.), Volcanism, impacts, and mass extinctions: Causes and effects (pp. 29–55). Geological Society of America. [Google Scholar]
- 232. Darroch, S. A. F. , Smith, E. F. , Laflamme, M. , & Erwin, D. H. (2018). Ediacaran extinction and Cambrian explosion. Trends in Ecology & Evolution, 33, 653–663. [DOI] [PubMed] [Google Scholar]
- 233. Schrag, D. P. , Berner, R. A. , Hoffman, P. F. , & Halverson, G. P. (2002). On the initiation of a snowball Earth. Geochemistry, Geophysics, Geosystems, 3, 1–21. [Google Scholar]
- 234. Goddéris, Y. , Donnadieu, Y. , Nédélec, A. , Dupré, B. , Dessert, C. , Grard, A. , Ramstein, G. , & François, L. M. (2003). The Sturtian ‘snowball’ glaciation: Fire and ice. Earth and Planetary Science Letters, 211, 1–12. [Google Scholar]
- 235. Cox, G. M. , Halverson, G. P. , Stevenson, R. K. , Vokaty, M. , Poirier, A. , Kunzmann, M. , Li, Z.‐X. , Denyszyn, S. W. , Strauss, J. V. , & Macdonald, F. A. (2016). Continental flood basalt weathering as a trigger for Neoproterozoic Snowball Earth. Earth and Planetary Science Letters, 446, 89–99. [Google Scholar]
- 236. Tabor, C. R. , Feng, R. , & Otto‐Bliesner, B. L. (2019). Climate responses to the splitting of a supercontinent: Implications for the breakup of Pangea. Geophysical Research Letters, 46, 6059–6068. [Google Scholar]
- 237. Foley, B. J. , & Driscoll, P. E. (2016). Whole planet coupling between climate, mantle, and core: Implications for rocky planet evolution. Geochemistry, Geophysics, Geosystems, 17, 1885–1914. [Google Scholar]
- 238. Cordani, U. G. , D'agrella‐Filho, M. S. , Brito‐Neves, B. B. , & Trindade, R. I. F. (2003). Tearing up Rodinia: The Neoproterozoic palaeogeography of South American cratonic fragments. Terra Nova, 15, 350–359. [Google Scholar]
- 239. Rainbird, R. , Cawood, P. A. , & Gehrels, G. (2012). The great Grenvillian sedimentation episode: Record of supercontinent Rodinia's assembly. In Busby, C. , & Azor, A. (Eds.), Tectonics of sedimentary basins: Recent advances (pp. 585–601). Chichester: Wiley‐Blackwell. [Google Scholar]
- 240. Slabunov, A. I. , Guo, J. , Balagansky, V. V. , Lubnina, N. V. , & Zhang, L. (2017). Early Precambrian crustal evolution of the Belomorian and Trans‐North China orogens and supercontinents reconstruction. Geodynamics and Tectonophysics, 8, 569–572. [Google Scholar]
- 241. Royer, D. L. , Berner, R. A. , Montañez, I. P. , Neil, J. T. , & Beerling, D. J. (2004). CO2 as a primary driver of Phanerozoic climate. GSA Today, 14, 4–10. [Google Scholar]