Hostname: page-component-848d4c4894-5nwft Total loading time: 0 Render date: 2024-05-19T18:35:24.840Z Has data issue: false hasContentIssue false

Transgression–regression cycles drive correlations in Ediacaran–Cambrian rock and fossil records

Published online by Cambridge University Press:  04 December 2023

Daniel C. Segessenman*
Affiliation:
Department of Geoscience, University of Wisconsin–Madison, Madison, Wisconsin 53706, U.S.A.
Shanan E. Peters
Affiliation:
Department of Geoscience, University of Wisconsin–Madison, Madison, Wisconsin 53706, U.S.A.
*
Corresponding author: Daniel C. Segessenman; Email: dsegesse@gmu.edu

Abstract

Strata of the Ediacaran Period (635–538.8 Ma) yield the oldest known fossils of complex, macroscopic organisms in the geologic record. These “Ediacaran-type” macrofossils (known as the Ediacaran biota) first appear in mid-Ediacaran strata, experience an apparent decline through the terminal Ediacaran, and directly precede the Cambrian (538.8–485.4 Ma) radiation of animals. Existing hypotheses for the origin and demise of the Ediacaran biota include: changing oceanic redox states, biotic replacement by succeeding Cambrian-type fauna, and mass extinction driven by environmental change. Few studies frame trends in Ediacaran and Cambrian macroevolution from the perspective of the sedimentary rock record, despite well-documented Phanerozoic covariation of macroevolutionary patterns and sedimentary rock quantity. Here we present a quantitative analysis of North American Ediacaran–Cambrian rock and fossil records from Macrostrat and the Paleobiology Database. Marine sedimentary rock quantity increases nearly monotonically and by more than a factor of five from the latest Ediacaran to the late Cambrian. Ediacaran–Cambrian fossil quantities exhibit a comparable trajectory and have strong (rs > 0.8) positive correlations with marine sedimentary area and volume flux at multiple temporal resolutions. Even so, Ediacaran fossil quantities are dramatically reduced in comparison to the Cambrian when normalized by the quantity of preserved marine rock. Although aspects of these results are consistent with the expectations of a simple fossil preservation–induced sampling bias, together they suggest that transgression–regression and a large expansion of marine shelf environments coincided with the diversification of animals during a dramatic transition that is starkly evident in both the sedimentary rock and fossil records.

Type
Article
Creative Commons
Creative Common License - CCCreative Common License - BY
This is an Open Access article, distributed under the terms of the Creative Commons Attribution licence (http://creativecommons.org/licenses/by/4.0/), which permits unrestricted re-use, distribution and reproduction, provided the original article is properly cited.
Copyright
Copyright © The Author(s), 2023. Published by Cambridge University Press on behalf of The Paleontological Society

Non-technical Summary

Ediacaran-age sedimentary rocks (635–538.8 million years ago) contain the oldest animal fossils that are visible to the naked eye. Several explanations have been suggested for the origins of animals in the Ediacaran, their disappearance at the end of the Ediacaran, and the following Cambrian explosion of animals (538.8–485.4 million years ago). For this study, we examined Ediacaran–Cambrian evolutionary patterns and how fossils (data from the Paleobiology Database) are related to the amount of sedimentary rock (data from Macrostrat) from the same time. Amounts of Cambrian rock increase to more than five times the amount of rock in the Ediacaran. The number of fossils increases in an equally dramatic manner from the Ediacaran to the Cambrian, and there are strong positive correlations between the amount of rock and the number of fossils. It is well known that in the Cambrian, sea level rose, leading to the flooding of the North American continent. This relative rise in sea level would have increased the amount of rock deposited on the continent. Cambrian flooding of the continent would have also provided a wider variety of shallow-marine environments for Cambrian animals to expand into, providing at least a partial explanation for the dramatic increase in the number and physical diversity of Cambrian fossils. A smaller flooding event during the Ediacaran may have enabled early fossil animals to develop evolutionary traits for shallow-marine environments that allowed them to rapidly evolve during the larger flooding in the Cambrian. The results of this study demonstrate that relative sea-level rise and associated continental-scale flooding known to influence the amount of rock may have played a role in shaping evolutionary patterns of Earth's earliest animals.

Introduction

The oldest complex macrofossils are found globally in sedimentary rocks of Ediacaran age (Sprigg Reference Sprigg1947; Glaessner Reference Glaessner1959; Knoll and Carroll Reference Knoll and Carroll1999; Xiao and Laflamme Reference Xiao and Laflamme2009; Xiao and Narbonne Reference Xiao, Narbonne, Gradstein, Ogg, Schmitz and Ogg2020). These distinctive, phylogenetically enigmatic fossils, often referred to as “Ediacaran-type macrofossils” or the “Ediacaran biota,” include taxa that are recognized as the oldest known metazoans (Droser and Gehling Reference Droser and Gehling2015; Bobrovskiy et al. Reference Bobrovskiy, Hope, Ivantsov, Nettersheim, Hallmann and Brocks2018a; Muscente et al. Reference Muscente, Bykova, Boag, Buatois, Mángano, Eleish and Prabhu2019; Wood et al. Reference Wood, Liu, Bowyer, Wilby, Dunn, Kenchington, Cuthill, Mitchell and Penny2019; Evans et al. Reference Evans, Hughes, Gehling and Droser2020; Dunn et al. Reference Dunn, Liu, Grazhdankin, Vixseboxse, Flannery-Sutherland, Green, Harris, Wilby and Donoghue2021; Shore et al. Reference Shore, Wood, Butler, Zhuravlev, McMahon, Curtis and Bowyer2021) and taxa now recognized as non-metazoan (e.g., Beltanelliformis; Bobrovskiy et al. Reference Bobrovskiy, Hope, Krasnova, Ivantsov and Brocks2018b). Since the Ediacaran's addition to the Geologic Time Scale (Knoll et al. Reference Knoll, Walter, Narbonne and Christie-Blick2006), significant advancements have been made in correlating its fossil-bearing stratigraphy, resulting in a general global division between a pre-Gaskiers (Pu et al., Reference Pu, Bowring, Ramezani, Myrow, Raub, Landing, Mills, Hodgin and Macdonald2016) lower Ediacaran sequence typically dominated by microfossil assemblages (but not limited to them; e.g., Liu and Tindal, Reference Liu and Tindal2021; Yang et al., Reference Yang, Li, Selby, Wan, Guan, Zhou and Li2022) and an upper Ediacaran post-Gaskiers sequence that bears the Ediacaran biota (e.g., Rooney et al. Reference Rooney, Cantine, Bergmann, Gómez-Pérez, Al Baloushi, Boag, Busch, Sperling and Strauss2020; Xiao and Narbonne Reference Xiao, Narbonne, Gradstein, Ogg, Schmitz and Ogg2020; Yang et al. Reference Yang, Rooney, Condon, Li, Grazhdankin, Bowyer, Hu, Macdonald and Zhu2021). The Ediacaran biota typically disappears by the Ediacaran/Cambrian transition (particularly in North America) and gives way to the distinctive faunal assemblages of the early Cambrian (Darroch et al. Reference Darroch, Smith, Laflamme and Erwin2018; Muscente et al. Reference Muscente, Bykova, Boag, Buatois, Mángano, Eleish and Prabhu2019; Bowyer et al. Reference Bowyer, Zhuravlev, Wood, Shields, Zhou, Curtis, Poulton, Condon, Yang and Zhu2022). There are many hypotheses concerning the appearance and disappearance of the Ediacaran biota, including changing redox (oxic/anoxic) states in Ediacaran oceans (Sperling et al. Reference Sperling, Carbone, Strauss, Johnston, Narbonne and Macdonald2016; Zhang et al. Reference Zhang, Xiao, Romaniello, Hardisty, Li, Melezhik and Pokrovsky2019), preservational biases caused by unique Ediacaran taphonomy or lack of outcrop (Seilacher Reference Seilacher, Holland and Trendall1984; Laflamme et al. Reference Laflamme, Darroch, Tweedt, Peterson and Erwin2013; Gehling et al. Reference Gehling, García-Bellido, Droser, Tarhan and Runnegar2019; Cuthill Reference Cuthill2022), and environmental catastrophe or biotic replacement–driven mass extinction (Darroch et al. Reference Darroch, Smith, Laflamme and Erwin2018; Tarhan et al. Reference Tarhan, Droser, Cole and Gehling2018; Zhang et al. Reference Zhang, Chen, Huang, Liu, Huang, Yeasmin and Fu2021). All of these proposed hypotheses invoke mechanisms known to exert controls on macroevolutionary trends observed from marine metazoan fossils in the Phanerozoic (Valentine Reference Valentine1969; Raup and Sepkoski Reference Raup and Sepkoski1982; Stanley Reference Stanley2007; Erwin Reference Erwin2008; Alroy Reference Alroy2010; Hannisdal and Peters Reference Hannisdal and Peters2011; Aberhan and Kiessling Reference Aberhan, Kiessling and Talent2012; and many others). However, few studies have examined the relationship between preserved rock quantity and macroevolution during the Ediacaran and across the Ediacaran/Cambrian transition.

Correlation of macroevolutionary patterns and sedimentary rock volume in deep time is a well-documented phenomenon (Newell Reference Newell1959; Raup Reference Raup1972, Reference Raup1976; Sepkoski et al. Reference Sepkoski, Bambach, Raup and Valentine1981; Peters and Foote Reference Peters and Foote2001, Reference Peters and Foote2002; Smith Reference Smith2001, Reference Smith2007; Smith et al. Reference Smith, Gale and Monks2001; Peters Reference Peters2005, Reference Peters2006; Smith and McGowan Reference Smith and McGowan2007; McGowan and Smith Reference McGowan and Smith2008; Heim and Peters Reference Heim and Peters2011; Peters et al. Reference Peters, Kelly and Fraass2013; Rook et al. Reference Rook, Heim and Marcot2013; Dunhill et al. Reference Dunhill, Hannisdal and Benton2014; Benton Reference Benton2015; Benson et al. Reference Benson, Butler, Close, Saupe and Rabosky2021). Sloss sequences linked to expansion and contraction of marine shelf area have also long been recognized in Phanerozoic strata as a second-order (~107 yr) control on the continental distribution of sedimentary rocks (e.g., Sloss Reference Sloss1963; Mackenzie and Pigott Reference Mackenzie and Pigott1981; Haq et al. Reference Haq, Hardenbol and Vail1987; Miller et al. Reference Miller, Kominz, Browning, Wright, Mountain, Katz, Sugarman, Cramer, Christie-Blick and Pekar2005; Haq and Schutter Reference Haq and Schutter2008; Meyers and Peters Reference Meyers and Peters2011; Peters and Heim Reference Peters and Heim2011b; Nance et al., Reference Nance, Murphy and Santosh2014; Husson and Peters Reference Husson and Peters2018). A matter of continuing debate is whether or not the correlation between rock and fossil records in deep time is indicative of preservation bias distorting patterns observed in the fossil record, or whether it is instead a signal of geologic process that acted as a “common cause” mechanism, driving both patterns of biological diversity and preserved rock quantity (Crampton et al. Reference Crampton, Beu, Cooper, Jones, Marshall and Maxwell2003; Peters Reference Peters2008; Peters and Heim Reference Peters, Heim, McGowan and Smith2011a; Peters et al. Reference Peters, Kelly and Fraass2013, Reference Peters, Quinn, Husson and Gaines2022; Holland Reference Holland2017; Husson and Peters Reference Husson and Peters2018; Nawrot et al. Reference Nawrot, Scarponi, Azzarone, Dexter, Kusnerik, Wittmer, Amorosi and Kowalewski2018). A Sloss sequence–like signal (“Mackenzie” sequence) coinciding with an increase in the number of rock units containing Ediacaran macrofossils has been observed in a macrostratigraphic analysis of a new compilation for the Ediacaran System in North America (Segessenman and Peters Reference Segessenman, Peters, Whitmeyer, Williams, Kellett and Tikoff2023). A more detailed analysis examining relationships between the rock and fossil records in this new compilation is warranted to provide new perspective on the relationship between macroevolutionary trends and sedimentary patterns during the Ediacaran/Cambrian transition. Here we present a quantitative analysis of intersecting rock and fossil datasets from the data platforms of Macrostrat (https://macrostrat.org; Peters et al. Reference Peters, Husson and Czaplewski2018; Segessenman and Peters Reference Segessenman, Peters, Whitmeyer, Williams, Kellett and Tikoff2023) and the Paleobiology Database (PBDB; https://paleobiodb.org).

Methods

Boundary ages, thicknesses, and lithologies of 546 revised Ediacaran (Segessenman and Peters Reference Segessenman, Peters, Whitmeyer, Williams, Kellett and Tikoff2023) and 2063 Cambrian (Peters et al. Reference Peters, Husson and Czaplewski2018) Macrostrat rock units (https://macrostrat.org) were matched (by location and rock unit name) to 412 Ediacaran and 16,133 Cambrian North American fossil occurrences from the PBDB (https://paleobiodb.org), accessed using the PBDB application programming interface (Peters and McClennen Reference Peters and McClennen2016). Rock unit age models and characteristics were compiled and established as part of previous studies (Peters et al. Reference Peters, Husson and Czaplewski2018; Segessenman and Peters Reference Segessenman, Peters, Whitmeyer, Williams, Kellett and Tikoff2023: p. 401). The rock unit age models were not adjusted for this study; instead, only the maximum and minimum ages of PBDB fossil occurrences were modified to reflect the constraints of the stratigraphic age models. Fossil occurrences are a fundamental unit in the PBDB and are defined as an instance of a particular organism at a particular location in time and space. PBDB fossil occurrences that did not have an exact matching unit name in the Macrostrat dataset were assigned to the Macrostrat unit that was geographically nearest and temporally overlapping and that shared a lithology with the collection-listed lithology. A total of 1088 Ediacaran and Cambrian PBDB occurrences that did not have any taxonomic information or that had a match distance (between PBDB occurrence coordinates and Macrostrat column centroids) greater than 300 km and no direct Macrostrat unit name match were removed from the dataset. An additional 20 PBDB occurrences with low-resolution age assignments such as “Neoproterozoic” (e.g., Grypania spiralis) were also removed. Ichnofossil occurrences were not removed but were restricted to calculations of fossil occupancy in the rock record; that is, a rock unit would be counted as “occupied” by fossils if it contained at least one fossil (including ichnofossil) occurrence. With the aforementioned approach, 403 Ediacaran and 15,034 Cambrian fossil occurrences were matched to 40 Ediacaran and 322 Cambrian Macrostrat rock units at the stratigraphic levels of member, formation, or group. Raw tables of Macrostrat units matched to PBDB occurrences and the R scripts used to generate figures/tables for this study can be found in the Supplementary Material (Supplement S1—Code). A table of Macrostrat unit_id's, stratigraphic names, counts of PBDB occurrences matched to each stratigraphic name, and modeled unit bounding ages is available in Supplementary Table S1. In addition, a table of PBDB fossil occurrences, their PBDB assigned stratigraphies/ages, and the Macrostrat matched stratigraphies/ages is available in Supplementary Table S2.

For each fossil occurrence, the PBDB reported minimum and maximum ages were used, unless the occurrence ages exceeded the modeled boundary ages of the containing rock unit (for further descriptions of Macrostrat boundary age models, see Peters et al. Reference Peters, Husson and Czaplewski2018; Segessenman and Peters Reference Segessenman, Peters, Whitmeyer, Williams, Kellett and Tikoff2023). For example, an Ediacaran PBDB occurrence with a minimum age of 538.8 Ma and a maximum age of 635 Ma (most Ediacaran PBDB entries have these assigned ages) matched to a rock unit with an upper boundary age of 550 Ma and a lower boundary age of 580 Ma would be given a new, narrower age range of 580–550 Ma. In this way, PBDB occurrence minimum and maximum ages were bound to their matched rock units’ bounding ages within Macrostrat's continuous time age model. The number of genera were derived from the number of occurrences by counting the number of unique genus names among occurrences for each time interval. Rock units with sedimentary lithologies and a thickness of zero (no available published thickness estimate) were given a median thickness calculated from thicknesses of the 10 most proximal (within 250 km), temporally overlapping rock units with sedimentary lithologies. This resulted in simulated thicknesses (median of all simulated thicknesses was ~358 m) for 151 Cambrian age units, which comprises ~7% of all North American Cambrian rock units. All calculations of rock and fossil quantities were made with 1 Myr time steps (consistent with a “continuous” age model construction), except for correlations between rock and fossil quantities, which were computed with 1, 5, and 10 Myr bins. The 10 Myr bins are the primary focus of our correlation analysis, as that time span is the resolution most appropriate for second-order scale influences (107 yr) and for the resolution of geochronologic constraints in the Ediacaran and early Cambrian.

Once all relevant Ediacaran–Cambrian rock unit and fossil occurrence characteristics were matched by formation name and/or spatiotemporal overlap, aggregate metrics were computed. Metrics used to describe the PBDB fossil dataset include counts of fossil occurrences; genus richness; and Shannon H indices for diversity, occurrences among lithologies, and occurrences among locations (Shannon Reference Shannon1948). Fossil occurrence locations were identified using the PBDB “states” field, which records the state or province in which the occurrence is located. Shannon H values (Shannon-Wiener Index) were generated for each 1 Myr time step (635–485 Ma) for both occurrences among genera and occurrences among geographic locations using the diversity function from the R package vegan (Oksanen et al. Reference Oksanen, Simpson, Blanchet, Kindt, Legendre, Minchin and O'Hara2022). Metrics used to describe the Macrostrat marine sedimentary rock datasets include counts of rock units, preserved area (km2), volume flux (km3/Myr), median rock unit thickness (m), and median unit duration (Myr). Bootstrap resampling (“block” sampling method with a 7 Myr moving window) was used to generate 2σ confidence intervals for the number of fossil occurrences, genera, counts of sedimentary units (with and without occurrences), median rock unit thickness, and median rock unit duration.

Counts of occupied rock (rocks that contain at least one occurrence), Spearman rank correlation coefficients of time-series first differences, and Spearman rank correlation coefficients (r s) of raw rock and fossil metrics were calculated to describe the intersection of the rock and fossil datasets. Correlation calculations for the Ediacaran dataset were temporally limited to 585 Ma and younger due to a lack of fossil data pre–585 Ma. Sedimentary rock lithologies were grouped into two general categories for this study: siliciclastics and carbonates. Macrostrat rock units include lithology as a relative percentage (e.g., 70% limestone, 15% sandstone, 15% shale) and general depositional environment (e.g., marine, nonmarine). Only marine sedimentary rock proportions that fit within the two general lithologic categories were included in calculations and time series for this study. To facilitate more direct comparison of rock quantities from the mesostrat Ediacaran dataset and the “whole-crust” Cambrian Macrostrat dataset, Ediacaran mesostrat column areas were scaled by a factor of 1.85; the justification for this is that the two compilations used different methods to determine column geographic footprints, which leads to a scalar offset in the area estimate for the same body of rock (for scalar calculation and discussion, see Segessenman and Peters Reference Segessenman, Peters, Whitmeyer, Williams, Kellett and Tikoff2023: p. 405).

Results

Trends in Ediacaran–Cambrian Rock and Fossil Records

Maps of rock and fossil locations were plotted to show the geographic distribution of marine sedimentary rock–bearing stratigraphic columns and fossil collections (a set of PBDB occurrences that are collocated geographically and temporally) across North America grouped by subdivisions of Ediacaran and Cambrian time (Fig. 1). Fossil collections are generally more widespread at times with increased marine sedimentary rock (represented by number of columns) and are less widespread at times with decreased marine sedimentary rock (Fig. 1). An animation of rock and fossil locations on North America in 5 Myr bins is available as Supplementary Figure S1. Note that Macrostrat columns include subsurface data, and therefore sedimentary rock is generally much more widespread than fossil collections, which are restricted primarily to outcrop belts.

Figure 1. Maps of North America with sediment-bearing column areas from Macrostrat (colored polygons) and fossil collection locations from the Paleobiology Database (PBDB; blue diamonds). Fossil collection locations have been randomly offset by a factor of 0.5%. Total numbers of columns and fossil collections are shown on each map. A, Lower Ediacaran (635–590 Ma); B, upper Ediacaran (590–538.8 Ma); C, Terreneuvian (538.8–521 Ma); D, Series 2 (521–509 Ma); E, Miaolingian (509–497 Ma); F, Furongian (497–485.4 Ma). Lower Ediacaran and upper Ediacaran informal divisions based on initial rise in preserved sediment area and volume. Cambrian epoch timings based on Cohen et al. (Reference Cohen, Finney, Gibbard and Fan2013; updated v2022/10).

Starting at ca. 585 Ma, the number of fossil occurrences and genera increase to a late Ediacaran maximum, chiefly due to Mistaken Point collections, (ca. 565 Ma; >130 occurrences) then sharply decrease and plateau until the Ediacaran/Cambrian transition (Fig. 2A). Fossil occurrences remain (locally) decreased across the Ediacaran/Cambrian boundary, but monotonically increase through the early and mid-Cambrian to a maximum (>3200 occurrences) after ca. 515 Ma (Fig. 2A). The maximum number of occurrences and genera are both an order of magnitude greater in the Cambrian than in the Ediacaran, with trends in the number of genera generally following that of occurrences (Fig. 2A). Increasing numbers of occurrences and genera coincide with increasing marine sedimentary unit counts, area, and volume flux in the Ediacaran and Cambrian (Fig. 2A–D). Rock–fossil fluctuations broadly correspond to the Mackenzie and Sauk Sloss sequences, apart from three deviations: (1) the ca. 570 Ma increase in the number of occurrences and genera does not coincide with an increase in sedimentary unit counts, rock area, or volume flux; (2) an Ediacaran volume flux increase to a period maximum (ca. 550 Ma) coincides with a decrease and local plateau in the number of fossil occurrences and genera; and (3) late Cambrian fossil occurrences/genera experience little change during a period maximum in rock area and significantly decreased volume flux (Fig. 2A–D).

Figure 2. Time series of rock and fossil metrics from the Ediacaran and Cambrian with Mackenzie and Sauk Sloss sequences and the Ediacaran/Cambrian transition highlighted. A, Log-scale plot of the number of occurrences and genera with bootstrap resampling–generated confidence intervals (2σ); B, log-scale plot of the total number of sedimentary rock units and the number of rock units that contain at least one fossil occurrence with bootstrap resampling–generated confidence intervals (2σ); C, stacked area plot of preserved marine sedimentary rock area (km2) divided into clastic (grain size–based) and carbonate categories; D, stacked area plot of calculated marine sedimentary volume flux (km3/Myr); and E, proportion of fossil-occupied sedimentary units (black), area (blue), and volume flux (red). Note that pre–580 Ma occurrences include fossil data from thicker, undivided stratigraphic sections with few geochronologic constraints.

The proportion of occupied marine sedimentary rock area (area of rock units that contain at least one occurrence, including ichnofossils) remains at or below 20%, and unit counts remain at or below 10% after 580 Ma in the Ediacaran (Fig. 2E). Occurrence occupancy falls to nearly 5% by the latest Ediacaran before increasing through the early and mid-Cambrian to a local maximum of nearly 25% (ca. 515 Ma), despite the fact that much of the Cambrian rock record is in the subsurface of North America. The occupied proportion of volume flux fluctuates greatly during the Ediacaran because of greater sensitivity to overall lower preserved rock volume and fossil occurrences. For example, the large pulse to nearly 60% of volume occupancy is due to a small number of late Ediacaran occurrences in thick, undivided stratigraphic sections in the SE United States (Segessenman and Peters Reference Segessenman, Peters, Whitmeyer, Williams, Kellett and Tikoff2023; Fig. 1E). Cambrian proportions of occupied sedimentary volume fluctuate to a much lesser degree due to a greater quantity of preserved rock and better geochronologic constraints (Fig. 2E). These results highlight that the character of the sedimentary record changes dramatically across the Ediacaran/Cambrian boundary, providing a strong physical justification for the position of the system and the Proterozoic/Phanerozoic eon boundary. Additionally, the quantified rock and fossil records exhibit parallel changes that broadly correspond to the Mackenzie and Sauk Sloss sequences.

Character of Fossil-bearing Ediacaran–Cambrian Rock Units

Median thickness and duration of sedimentary rock units differ between occupied units and all sedimentary rock units during the Ediacaran (Fig. 3A,B). The median thickness of all sedimentary rock units is relatively high in the early Ediacaran (due to a relatively low number of columns with thick, undivided sections), declines after ca. 590 Ma to a mid- to late Ediacaran plateau, and decreases continuously after the Ediacaran/Cambrian transition (Fig. 3A). The median thickness of occupied sedimentary units fluctuates dramatically during the mid- to late Ediacaran, reaching highs greater than all sedimentary units at ca. 575–550 Ma before a local maximum during the Ediacaran/Cambrian transition (Fig. 3A). By ca. 530 Ma, the median thickness of occupied sedimentary units fluctuates much less and remains consistent with all sedimentary units as it generally decreases through the remainder of the Cambrian (Fig. 3A). The significant increases of occupied unit thickness at ca. 575–550 Ma are likely due to few geochronologic constraints and low total preserved sediment volume that results in thicker, undivided stratigraphic sections (Fig. 3A). However, the increase in median thickness at the Ediacaran/Cambrian transition may also be due to regression marking the end of the Mackenzie sequence, which would have left only the thickest and most continuous stratigraphic sections on the continental margins.

Figure 3. Time series of median thickness and duration of sedimentary rocks and the number of occurrences reported from clastic and carbonate lithologies with Mackenzie and Sauk Sloss sequences and the Ediacaran/Cambrian transition highlighted. A, Median thickness (m) of all sedimentary units (black) with bootstrap resampling–generated confidence interval (2σ) and only sedimentary units that are occupied (contain at least one occurrence; red); B, median duration (Myr) of all sedimentary units (black) with bootstrap resampling–generated confidence interval (2σ) and only sedimentary units that are occupied (red); and C, stacked area of occurrence counts by Paleobiology Database (PBDB) reported lithology. A single occurrence can have multiple lithologies and therefore can be counted within multiple lithologic categories for one time interval.

Median durations of all sedimentary units (black line) follow a similar trend to that of unit median thicknesses (Fig. 3B). Median duration is relatively high during the early Ediacaran; decreases to a plateau starting at ca. 585 Ma, except for an increase from 567 to 563 Ma; increases more significantly during the Ediacaran/Cambrian transition; and decreases for most of the Cambrian (Fig. 3B). The median durations of all sedimentary units and occupied sedimentary units are noticeably elevated during the Ediacaran/Cambrian transition, a departure from the median thickness (Fig. 3A,B). The median duration of occupied units from the mid- to late Ediacaran follows a similar trend to that of all sedimentary units, although it is lower relative to total sedimentary units, except during the Ediacaran/Cambrian transition (Fig. 3B). The Ediacaran/Cambrian transition increase in the median duration of occupied units is most likely due to very few fossil occurrences reported within the few thick, continuous sections of rock that span this boundary (Fig. 3B). In a similar manner to that of the median thickness, marine regression may have contributed to the median duration increase observed in all sedimentary units during the Ediacaran/Cambrian transition through the reduction of shorter-duration, more-proximal stratigraphic sections (Fig. 3A,B).

The lithologies yielding fossil occurrences exhibit distinct differences between the Ediacaran and Cambrian (Fig. 3C). The comparatively rare Ediacaran occurrences are almost exclusively reported from siliciclastic lithologies in contrast to early Cambrian occurrences, which are dominantly from carbonates (Fig. 3C). Interestingly, Ediacaran occurrences are reported primarily from siliciclastic lithologies, despite increased proportions of carbonate at the same time (ca. 577–555 Ma; Figs. 2C,D, and 3C). At ca. 515 Ma, there is a rapid increase in the proportion of Cambrian fossil occurrences reported from fine-grain sedimentary rocks (Fig. 3C), although fossils reported from carbonate lithologies dominate the majority of the Cambrian rock record (Fig. 2C,D). Overall, these results are consistent with known differences in the preservation of Ediacaran (largely preserved in microbial mat–influenced siliciclastics) and Cambrian (large increase in calcifiers) taxa that generally mirror changes in the nature of Ediacaran–Cambrian sedimentary units.

Ediacaran–Cambrian Macroevolutionary Trends

Fossil occurrences and unique genera were normalized by counts of sedimentary units, preserved rock area, and volume flux for the Ediacaran and Cambrian (Fig. 4). Even when the decreased sedimentary rock quantity in the Ediacaran is accounted for, the numbers of occurrences and unique genera in the Cambrian rapidly surpass those of the Ediacaran (Fig. 4). The late Ediacaran maximum in fossil occurrences and genera from ca. 570 to 555 Ma (Fig. 2A) remains the most significant increase in the Ediacaran when normalized by rock quantities (Fig. 4). Increases in normalized occurrences and genera during the Cambrian are present but muted (Fig. 4) when compared with the raw values (Fig. 2A). However, two intervals appear to be significant from this perspective: (1) the increase of normalized occurrences and genera from ca. 520 to 505 Ma and (2) the volume flux–normalized increase of occurrences and genera during the latest Cambrian after ca. 497 Ma (Fig. 4). The sharp increase in occurrences and genera from ca. 515 to 505 Ma is because of sampling from the Stephen Formation (includes the Burgess Shale), which makes up ~10% of Cambrian occurrences in this study's dataset (Fig. 4). The volume flux–normalized increase of occurrences and genera after ca. 497 Ma can be attributed to the decreased volume at the end of the Cambrian coinciding with little change in the number of occurrences and genera (Figs. 2A,D, and 4).

Figure 4. Time series of occurrence and genus counts normalized by sedimentary rock units (occurrences/rock unit), rock area (occurrences/10,000 km2), and volume flux (occurrences/1000 km3/Myr) with Mackenzie and Sauk Sloss sequences and the Ediacaran/Cambrian transition highlighted. A, Counts of occurrences normalized to sedimentary rock quantities; and B, counts of genera normalized to sedimentary rock quantities. See text for discussion.

The Shannon H index of generic diversity was calculated for each 1 Myr time step through the Ediacaran and Cambrian to summarize the distribution of occurrences among genera (Fig. 5A,B). Shannon H-values were also calculated to summarize the distribution of occurrences among locations to highlight periods of potentially uneven locality sampling (Fig. 5C). After the initial Ediacaran increase in the number of genera at ca. 585 Ma, generic Shannon H index values fluctuate, but remain close to a value of 3.5 until the latest Ediacaran (Fig. 5B). Cambrian Shannon H index values continually increase after the Ediacaran/Cambrian transition to a maximum of 6 by the latest Cambrian, indicating increasing generic diversity and “evenness” of occurrence frequency across genera (Fig. 5B). These results are interesting, because the ca. 570–555 Ma high in the number of occurrences and genera (Fig. 2A) does not stand out except for a minor pulse at 565 Ma (Fig. 5B). Furthermore, the Shannon H index calculated for the frequency of occurrences across location names during this same interval (ca. 570–555 Ma), exhibits a decrease that indicates a greater number of genera are reported from a less diverse pool of locations (Fig 5C). For this particular time period, the dominant source of occurrences (and therefore a source of sampling bias) is the Conception Group of Newfoundland (includes the Mistaken Point Fm.). This phenomenon is also present in the Cambrian record from ca. 515 to 505 Ma, the same interval in which samples from the Stephen Formation/Burgess Shale impact the normalized occurrence and genus curves (Figs. 4, 5C). Except for the aforementioned deviations and the Ediacaran/Cambrian transition, the overall diversity and evenness of sampling locations increases through the Ediacaran and Cambrian (Fig. 5C).

Figure 5. Time series of raw genus counts and the Shannon H indices of unique genus names and their reported locations (“states” field in the Paleobiology Database [PBDB]) with Mackenzie and Sauk Sloss sequences and the Ediacaran/Cambrian transition highlighted. A, Raw genus counts; B, Shannon H index of unique genera names and their occurrence frequencies; and C, Shannon H index of unique state names and their occurrence frequencies. Decreases in Shannon H indices from ca. 565–555 Ma and ca. 515–505 Ma represent intervals in which sampling is dominated by collections at specific localities (Mistaken Point Fm., Newfoundland, Canada, and Stephen Fm., British Columbia, Canada, respectively).

Correlations of Rock and Fossil Record

Spearman rank-order correlation coefficients (r s) and respective p-values for various metrics describing the rock and fossil records were calculated using 1, 5, and 10 Myr bins (Table 1). To calculate these correlations, the surrogateCor function from the astrochron package in R was used (Meyers Reference Meyers2014). The surrogateCor function was designed to calculate correlations and estimate the statistical significance of those correlations using the method of Ebisuzaki (Reference Ebisuzaki1997), in which time series are derived from stratigraphic successions. Correlations within a 95% or 90% confidence interval (p-value < 0.05 or 0.1, respectively) were considered (green and yellow shaded cells in Table 1, respectively). There is a very strong positive correlation between the number of occurrences and genera at all temporal resolutions (r s = 0.986, 0.951, 0.973; Table 1), making the two metrics largely interchangeable. The numbers of occurrences and genera are also both positively correlated with sedimentary area, although the number of genera exhibits a stronger positive correlation at the 5 and 10 Myr temporal resolutions (r s = 0.851, 0.891 and 0.856, 0.955, respectively; Table 1). A similar result is obtained when comparing occurrence and genus counts with the median unit thickness and carbonate flux. In all other cases, occurrences and genera have similar correlations or occurrences have slightly stronger correlations than genera (Table 1). Spearman correlation coefficients for the first differences of rock and fossil metrics were also calculated using 1, 5, and 10 Myr bins. The 10 Myr resolution first differences for number of sedimentary marine units, median unit thickness, total area, siliciclastic area (although only for occurrences), total volume flux, and carbonate volume flux had significant correlations (strong to moderate) with fossil metrics. The 5 Myr resolution first differences had less consistently significant correlations (moderate to weak), and 1 Myr resolution first differences had almost no significant correlations (Supplementary Table S3).

Table 1. Spearman rank correlation coefficients, ρ (r s), and associated p-values calculated between Ediacaran–Cambrian sedimentary rock and fossil quantities in 1, 5, and 10 Myr bins. Green cells represent correlation coefficients with corresponding p-values in the 95% confidence interval (p < 0.05), and yellow cells represent the 90% confidence interval (p < 0.1). Bold values are Spearman's rho (ρ) rank correlation coefficients. “No. of sed. marine units” is the number of sedimentary marine units.

A strong negative correlation is present between the number of genera/occurrences and the median duration and thickness of marine sedimentary units, possibly indicating that the presence of fossils leads to a greater potential for temporal subdivision of rock units into thinner intervals. Carbonate volume flux has a positive correlation with the number of fossil occurrences and genera at all resolutions, although only at the 90% confidence level (Table 1). There is no statistically significant correlation between the number of genera or fossil occurrences and the area of carbonates, except at the 10 Myr resolution for the number of fossil occurrences (Table 1). Correlations between sedimentary rock quantities and the fossil record were also calculated separately for the Ediacaran (post–585 Ma) and Cambrian periods (Supplementary Tables S4, S5). Post–585 Ma Ediacaran rock–fossil correlations are rarely statistically significant, due primarily to fewer data points and/or greater uncertainties in Ediacaran rock and fossil ages (Supplementary Table S4). Cambrian rock–fossil correlations show similar but weaker correlations to those of the combined Ediacaran–Cambrian rock–fossil records, with the notable exception of stronger correlations between fossil quantities and carbonate area/volume (Supplementary Table S5).

Discussion

Geochronologic and biostratigraphic controls are less resolved in the Ediacaran and early Cambrian than in much of the rest of the Phanerozoic. Although studies providing and refining taxonomy, biostratigraphy, chemostratigraphy, and radioisotopic dates from the Ediacaran system have increased in frequency globally since its formal addition to the Geologic Time Scale in 2006 (Knoll et al. Reference Knoll, Walter, Narbonne and Christie-Blick2006), it remains a formidable challenge to correlate Ediacaran sections regionally and globally. The age models used in this study for the Ediacaran and Cambrian systems of North America (for full description, see methods section of Segessenman and Peters Reference Segessenman, Peters, Whitmeyer, Williams, Kellett and Tikoff2023: pp. 401–403) were compiled with the intention of reflecting current published interpretations; a “state of the Ediacaran–Cambrian of North America.” We do not assert that our compilation and age models are without error, but that we have characterized the aggregate understanding of these systems in a stratigraphically self-consistent way, such that the overarching temporal trends will likely endure, even when new discoveries and analyses require the age models to be revised, expanding and/or compressing the stratigraphically grounded temporal patterns documented here. We present correlation results at the 1, 5, and 10 Myr resolutions, but the following discussion focuses primarily on the 10 Myr resolution results, as that is the resolution that is likely to be most reflective of the age model's precision.

There are strong positive correlations between raw time series of occurrences, diversity, and sedimentary rock quantities for the mid- to late Ediacaran and Cambrian (Table 1), but first differences are moderately to weakly correlated. The lack of strong significant correlation between first differences in rock and fossil metrics suggests that sampling bias is not a primary driver of the strong correlations that are evident in the raw metrics. This stands in contrast to the situation for most of the remaining Phanerozoic, where first differences in rock quantity and fossil occurrences/diversity are more strongly correlated (Crampton et al. Reference Crampton, Beu, Cooper, Jones, Marshall and Maxwell2003; Peters and Heim Reference Peters, Heim, McGowan and Smith2011a; Peters et al. Reference Peters, Kelly and Fraass2013), but the long-term trends in each diverge toward the Recent, with diversity continuing to increase and shallow-marine rock quantity remaining steady or even declining (Benson et al. Reference Benson, Butler, Close, Saupe and Rabosky2021; Peters et al. Reference Peters, Quinn, Husson and Gaines2022). The Ediacaran–Cambrian (635–485.4 Ma), by contrast, exhibits a significant decrease in rock quantity with increasing age (Fig. 2C,D), a pattern that is generally predicted by all models of erosion-dominated sedimentary rock cycling (Peters and Husson Reference Peters and Husson2017). In the case of the Ediacaran–Cambrian, though, it is apparent that the large decrease in sedimentary rock quantity with increasing age primarily reflects the signature of an increase in the depositional area of marine sediments throughout the late Ediacaran and Cambrian. This led to the progressive deposition of an increasingly expansive, relatively thin veneer (at least across the North American continental interior) of Cambrian marine sediment over area-limited Ediacaran sediments and a much wider area of exposed heterogeneous Precambrian igneous and metamorphic basement rocks (Peters and Gaines Reference Peters and Gaines2012). Regardless of whether these basement rocks were exhumed during Snowball Earth glaciations (Keller et al. Reference Keller, Husson, Mitchell, Bottke, Gernon, Boehnke, Bell, Swanson-Hysell and Peters2019; McDannell and Keller Reference McDannell and Keller2022) or during a more protracted, multistaged tectonic uplift history (e.g., Flowers et al. Reference Flowers, Macdonald, Siddoway and Havranek2020; Sturrock et al. Reference Sturrock, Flowers and Macdonald2021), it is clear that the Great Unconformity in North America is defined in large part by a shift from net continental denudation to net burial by Cambrian and younger sedimentary cover. This Phanerozoic cover has survived to the present day largely intact, although some unknown amount of Cambrian sediment has been lost from the Canadian Shield, thereby reducing the apparent increase in shelf area implied by the surviving record. Focused erosion sometime between the Ediacaran and Cambrian of a type not repeated in the later Phanerozoic seems unlikely to be driving the temporal trajectory of sedimentary rock quantity during this interval (Peak et al. Reference Peak, Flowers and Macdonald2023). Instead, continental-scale transgressive–regressive cycles are the probable drivers of observed trends in Ediacaran–Cambrian sedimentary rock quantities. Thus, the primary signal in the surviving sedimentary rock record is one of environmental change and real shifts in the extent of epicontinental marine sedimentation, not postdepositional modification of some markedly different environmental history.

Preserved sedimentary volume flux on continents is primarily controlled by accommodation and sediment supply (Miall Reference Miall2016). Evidence suggests that Laurentia, which constitutes the bulk of North America, had ample sediment supply during the Ediacaran but was generally accommodation limited due to an apparent lack of continental basins and limited continental flooding. Changes in accommodation would then have been driven primarily by local tectonics (such as that of Ediacaran–Cambrian Laurentian margin rifting; Macdonald et al. Reference Macdonald, Yonkee, Flowers, Swanson-Hysell, Whitmeyer, Williams, Kellett and Tikoff2023) and/or fluctuations in base level (as observed in “Western Laurentia” from Segessenman and Peters [2023]), either due to continental margin subsidence, global sea-level rise, or both. In light of this, an increase in preserved rock volume flux, and a more minor area increase, with a high proportion of carbonates after ca. 580 Ma, is interpreted as an increase in accommodation driven by base-level rise on the Laurentian margin. An increase in Ediacaran fossil occurrences and genera coincides with the post–580 Ma sedimentary volume flux increase and its subsequent decrease at the Ediacaran/Cambrian transition (Figs. 1–5; Table 1). Similarly, the dramatic radiation of organisms in the Cambrian is matched by an equally dramatic increase in the volume and area of sedimentary rock preserved on Laurentia, although the increase in Cambrian sedimentary volume cannot entirely explain the Cambrian's increased fossil occurrences and generic richness (Figs. 1, 2A–D).

Though correlation does not necessitate causation, it can be assumed that second-order (107 yr) changes in sedimentary area and volume flux are strongly influenced by changes in accommodation and are not influenced by changes in the number of fossil occurrences or genera. Preserved sedimentary volume can, however, influence the overall abundance of fossils and is subject to common cause mechanisms that can drive parallel changes in both the rock record and biological communities. Transgression, driven by subsidence and/or global sea-level rise, would have increased potential habitable ecospace, which in turn would create more potential environments in which organisms may be preserved (Fig. 1). This does not necessarily mean that the probability of preservation in a given environment increased, but it does imply that the number of organisms that could be preserved and recovered as fossils in North America increased. Although, the probability of preservation would have increased in the Cambrian due to the rapid diversification of calcifiers, which may have been enabled by (but not necessarily driven by; see Gilbert et al. Reference Gilbert, Bergmann, Boekelheide, Tambutté, Mass, Marin and Adkins2022) increased availability of carbonate-dominated shallow-marine environments (Knoll Reference Knoll2003; Fig. 3C). The combined effects of taphonomic change and ecospace expansion may help to explain the rapid, dramatic Cambrian increases in biodiversity, even when normalized to rock quantity (Fig. 4).

An increasing number of sedimentary units, particularly during the Cambrian, may represent increasing environmental heterogeneity and ecological opportunity (influencing macroevolution and taphonomy) as shallow-marine shelf habitat space expanded, potentially driving generic richness as well as an increase in the total number of organisms inhabiting an increasingly broad and heterogeneous shelf. Increasing environmental heterogeneity is also indicated by an observed increase in regional differences of faunal compositions coincident with the Cambrian radiation (Na et al. Reference Na, Kocsis, Li and Kiessling2022). This relationship may be evidenced by a stronger correlation between genera and preserved sedimentary area (0.955) than the number of occurrences and preserved sedimentary area (0.891), as well as by the fact that genera exhibit a strong correlation with the number of sedimentary units (Table 1). Regression at the end of the Ediacaran would have had the opposite effect and is evidenced by decreased sedimentary area and volume flux (Fig. 2C,D), an increase in the median duration of sedimentary units (Fig. 3B), and the presence of a globally occurring (though likely diachronous) sequence boundary across the Ediacaran/Cambrian transition (Shahkarami et al. Reference Shahkarami, Buatois, Mángano, Hagadorn and Almond2020; Bowyer et al. Reference Bowyer, Zhuravlev, Wood, Shields, Zhou, Curtis, Poulton, Condon, Yang and Zhu2022).

In addition to the stark changes in rock quantity and biodiversity discussed above, there is an equally dramatic shift in the overall character of metazoans from the Ediacaran to the Cambrian (Butterfield Reference Butterfield2009; Darroch et al. Reference Darroch, Smith, Laflamme and Erwin2018; Zhuravlev and Wood Reference Zhuravlev and Wood2018; Bowyer et al. Reference Bowyer, Zhuravlev, Wood, Shields, Zhou, Curtis, Poulton, Condon, Yang and Zhu2022). Although faunal compositions of the Ediacaran are clearly distinct from those of the Cambrian (Erwin Reference Erwin2021), morphologies and behaviors thought to originate in the Cambrian have been documented in late Ediacaran strata (Bengtson and Zhao Reference Bengtson and Zhao1992; Gehling and Droser Reference Gehling and Droser2018; Cai et al. Reference Cai, Xiao, Li and Hua2019; Wood et al. Reference Wood, Liu, Bowyer, Wilby, Dunn, Kenchington, Cuthill, Mitchell and Penny2019; Tarhan et al. Reference Tarhan, Myrow, Smith, Nelson and Sadler2020; Darroch et al. Reference Darroch, Cribb, Buatois, Germs, Kenchington, Smith and Mocke2021). However, Cambrian communities include an increasing number of calcifiers and taxa with larger maximum body sizes, and there are increased traces of more metabolically demanding behaviors such as complex feeding/burrowing patterns, increased motility, and increased predator–prey interactions (Schiffbauer et al. Reference Schiffbauer, Huntley, O'Neil, Darroch, Laflamme and Cai2016; Zhuravlev and Wood Reference Zhuravlev and Wood2020; Zhang and Shu Reference Zhang and Shu2021). Alongside significant environmental change discussed previously, two other major factors are cited as key drivers of Ediacaran–Cambrian metazoan macroevolution: (1) increasing atmospheric pO2 buildup that may have enabled the development of taxa with larger body sizes and more metabolically demanding behaviors (Och and Shields-Zhou Reference Och and Shields-Zhou2012; Lenton et al. Reference Lenton, Boyle, Poulton, Shields-Zhou and Butterfield2014; Chen et al. Reference Chen, Ling, Vance, Shields-Zhou, Zhu, Poulton and Och2015; He et al. Reference He, Zhu, Mills, Wynn, Zhuravlev, Tostevin, Pogge von Strandmann, Yang, Poulton and Shields2019; Cole et al. Reference Cole, Mills, Erwin, Sperling, Porter, Reinhard and Planavsky2020; Jiang et al. Reference Jiang, Zhao, Shen, Huang, Chen and Cai2022) and (2) significant geochemical change in shallow-marine environments, such as increased dissolved Ca2+ concentrations and availability of biolimiting nutrients that may have enabled increased prevalence of calcifying taxa (Brennan et al. Reference Brennan, Lowenstein and Horita2004; Peters and Gaines Reference Peters and Gaines2012; Wang et al. Reference Wang, Ling, Struck, Zhu, Zhu, He, Yang, Gamper and Shields2018; Li et al. Reference Li, Shi, Cheng, Jin and Algeo2020; Cherry et al. Reference Cherry, Gilleaudeau, Grazhdankin, Romaniello, Martin and Kaufman2022; Weldeghebriel et al. Reference Weldeghebriel, Lowenstein, García-Veigas and Cendón2022).

Atmospheric pO2 buildup through geologic time on Earth is directly related to increased burial of organic carbon (Berner Reference Berner1982), a process influenced by continental flooding shifting deposition from short-lived oceanic crust to long-lived continental reservoirs. Similarly, flooding of Laurentia during a time in which its surface geology may have largely consisted of exposed crystalline basement following Cryogenian glaciation has been cited as a potential source of increased biolimiting nutrients and Ca2+ concentrations in shallow-marine settings during the Cambrian. Cambrian continental flooding is an influential factor that, given the unique geologic and paleobiological contexts of the Ediacaran–Cambrian Earth, may have served as a driver of a “perfect storm” that enabled the Cambrian explosion of life, where minor flooding in the Ediacaran enabled metazoan biologic innovations that then truly “exploded” during the Cambrian Sauk transgression. Ultimately, the geologic process(es) driving the observed flooding signatures in the Ediacaran and Cambrian are matters of ongoing research, although mantle dynamics (Zou et al. Reference Zou, Mitchell, Chu, Brown, Jiang, Li, Zhao and Zhai2023), rift-related continental margin subsidence, and the locus of subduction globally (Macdonald et al. Reference Macdonald, Yonkee, Flowers, Swanson-Hysell, Whitmeyer, Williams, Kellett and Tikoff2023; Tasistro-Hart and Macdonald Reference Tasistro-Hart and Macdonald2023) have been cited as potential drivers.

The extent to which the results presented here are representative of global trends in Ediacaran–Cambrian macroevolution and macrostratigraphy has not been directly examined due to Macrostrat's current North American focus. However, it is recognized that early Ediacaran rock and fossil records are better preserved on other continents (e.g., China; Cunningham et al. Reference Cunningham, Vargas, Yin, Bengtson and Donoghue2017; Yang et al. Reference Yang, Li, Selby, Wan, Guan, Zhou and Li2022), and that fossiliferous latest Ediacaran to Ediacaran/Cambrian transition strata are more common on other continents (e.g., White Sea and Nama assemblages; Waggoner Reference Waggoner2003). Decreased rock volume at the latest Ediacaran on North America is consistent with regressive systems tracts identified at the Ediacaran/Cambrian transition globally (Bowyer et al. Reference Bowyer, Zhuravlev, Wood, Shields, Zhou, Curtis, Poulton, Condon, Yang and Zhu2022). However, the general lack of terminal Ediacaran biota fossils in North America could indicate that Laurentia's taxa were harder hit in an end-Ediacaran extinction event, that environmental conditions were particularly poor for preservation, that fossil-bearing strata have been eroded, or a combination of these. Early Ediacaran sections do exist on North America, but they are rarer and largely un-fossiliferous. This may be a result of conditions unfavorable to fossil preservation combined with low rock preservation but could also indicate that the Ediacaran biota did not originally develop in Laurentia, but arrived later. Despite these differences from global Ediacaran strata, our results are generally consistent with a mid- to late Ediacaran appearance of the Ediacaran biota, an apparent late Ediacaran diversity maximum, and, albeit earlier than global sections, an end-Ediacaran decline (Xiao and Narbonne Reference Xiao, Narbonne, Gradstein, Ogg, Schmitz and Ogg2020; Evans et al. Reference Evans, Tu, Rizzo, Surprenant, Boan, McCandless, Marshall, Xiao and Droser2022). In addition, our results, when combined with earlier Ediacaran fossils such as those of the Lantian or Weng'an biotas (Cunningham et al. Reference Cunningham, Vargas, Yin, Bengtson and Donoghue2017; Yang et al. Reference Yang, Li, Selby, Wan, Guan, Zhou and Li2022) and latest Ediacaran assemblages such as those of the White Sea or Nama (Waggoner Reference Waggoner2003), suggest a more protracted radiation of metazoans through the Ediacaran and a “less explosive” (but still greater magnitude and comparatively rapid) Cambrian radiation (Wood et al. Reference Wood, Liu, Bowyer, Wilby, Dunn, Kenchington, Cuthill, Mitchell and Penny2019; Servais et al. Reference Servais, Cascales-Miñana, Harper, Lefebvre, Munnecke, Wang and Zhang2023) that broadly mirrors patterns of continental transgression and regression recognized globally (Sloss Reference Sloss1963; Sears and Price Reference Sears and Price2003; Avigad et al. Reference Avigad, Sandler, Kolodner, Stern, McWilliams, Miller and Beyth2005; Lorentzen et al. Reference Lorentzen, Augustsson, Nystuen, Berndt, Jahren and Schovsbo2018).

We do not suggest that a mid-Ediacaran transgression (Mackenzie sequence) drove the origins of metazoans, that terminal Ediacaran regression functioned as a primary driver of Ediacaran-type fauna extinction, or that the radiation of life in the Cambrian was solely due to a coincident expansion of habitable shallow-shelf ecospace. Fossil occurrences and genera normalized to rock quantities indicate that sedimentary rock volume alone cannot explain all patterns in the fossil record (Fig. 4). Rather, the results presented herein provide new perspectives on transgression–regression cycles as strong environmental correlates of the appearance and diversification of the Ediacaran biota in the mid-Ediacaran, their apparent decline at the terminal Ediacaran, and the transition to (and rapid expansion of) Cambrian-type fauna during the Sauk transgression. Our results do not preclude any existing hypotheses driving evolution during the Ediacaran and Cambrian; instead, they demonstrate the influence of transgressive–regressive cycles in the observed sedimentary record at the dawn of animal life and provide a rock record–based framework within which to interpret and test existing hypotheses of macroevolutionary drivers. Expansion of the Macrostrat database to other continents and continued growth of the PBDB will enable further assessment of how the rock and fossil records covary at the dawn of animal life and the subsequent Cambrian explosion.

Acknowledgments

D.C.S. was supported by funding from the University of Wisconsin–Madison Geoscience Department, the Morgridge Distinguished Graduate Fellowship, and by the Atmospheric, Oceanic, and Earth Sciences Department at George Mason University. We would like to thank B. N. Hupp for their feedback on an initial draft of this article. We would also like to thank S. Evans and an anonymous reviewer for their detailed, constructive feedback that significantly improved this paper. Macrostrat infrastructure development was supported by U.S. National Science Foundation grant EAR-1150082 and EarthCube grant ICER-1440312. This is Paleobiology Database publication no. 466.

Competing Interests

The authors declare no competing interests.

Data Availability Statement

All supplementary data files for this study, including R scripts for analyses, an animation of fossil and stratigraphic column locations through time, tables of rock units matched to fossil occurrences, tables of fossil occurrence assigned ages, and correlations for Ediacaran and Cambrian rock and fossil quantities as separate time periods are available from the Dryad Digital Depository: https://doi.org/10.5061/dryad.xwdbrv1k9.

Footnotes

Present address: Atmospheric, Oceanic, and Earth Sciences Department, George Mason University, Fairfax, Virginia 22030, U.S.A.

References

Literature Cited

Aberhan, M., and Kiessling, W.. 2012. Phanerozoic marine biodiversity: a fresh look at data, methods, patterns and processes. Pp. 322 in Talent, J. A., ed. Earth and life. International Year of Planet Earth. Springer, Dordrecht.10.1007/978-90-481-3428-1_1CrossRefGoogle Scholar
Alroy, J. 2010. Geographical, environmental and intrinsic biotic controls on Phanerozoic marine diversification. Palaeontology 53:12111235.CrossRefGoogle Scholar
Avigad, D., Sandler, A., Kolodner, K., Stern, R. J., McWilliams, M., Miller, N., and Beyth, M.. 2005. Mass-production of Cambro-Ordovician quartz-rich sandstone as a consequence of chemical weathering of Pan-African terranes: environmental implications. Earth and Planetary Science Letters 240:818826.10.1016/j.epsl.2005.09.021CrossRefGoogle Scholar
Bengtson, S., and Zhao, Y.. 1992. Predatorial borings in late Precambrian mineralized exoskeletons. Science 257:367369.10.1126/science.257.5068.367CrossRefGoogle Scholar
Benson, R. B., Butler, R. J., Close, R, A., Saupe, E. E., and Rabosky, D. L.. 2021. Biodiversity across space and time in the fossil record. Current Biology 31:R1225R1236.10.1016/j.cub.2021.07.071CrossRefGoogle ScholarPubMed
Benton, M. J. 2015. Palaeodiversity and formation counts: redundancy or bias? Palaeontology 58:10031029.CrossRefGoogle Scholar
Berner, R. A. 1982. Burial of organic carbon and pyrite sulfur in the modern ocean: its geochemical and environmental significance. American Journal of Science 282:451473.10.2475/ajs.282.4.451CrossRefGoogle Scholar
Bobrovskiy, I., Hope, J. M., Ivantsov, A., Nettersheim, B. J., Hallmann, C., and Brocks, J.J.. 2018a. Ancient steroids establish the Ediacaran fossil Dickinsonia as one of the earliest animals. Science 361:12461249.CrossRefGoogle ScholarPubMed
Bobrovskiy, I., Hope, J. M., Krasnova, A., Ivantsov, A., and Brocks, J.J.. 2018b. Molecular fossils from organically preserved Ediacara biota reveal cyanobacterial origin for Beltanelliformis. Nature Ecology and Evolution 2:437440.CrossRefGoogle ScholarPubMed
Bowyer, F. T., Zhuravlev, A. Y., Wood, R., Shields, G. A., Zhou, Y., Curtis, A., Poulton, S. W., Condon, D. J., Yang, C., and Zhu, M.. 2022. Calibrating the temporal and spatial dynamics of the Ediacaran–Cambrian radiation of animals. Earth-Science Reviews 225:103913.10.1016/j.earscirev.2021.103913CrossRefGoogle Scholar
Brennan, S. T., Lowenstein, T. K., and Horita, J.. 2004. Seawater chemistry and the advent of biocalcification. Geology 32:473476.10.1130/G20251.1CrossRefGoogle Scholar
Butterfield, N. J. 2009. Macroevolutionary turnover through the Ediacaran transition: ecological and biogeochemical implications. Geological Society of London Special Publication 326:5566.10.1144/SP326.3CrossRefGoogle Scholar
Cai, Y., Xiao, S., Li, G., and Hua, H.. 2019. Diverse biomineralizing animals in the terminal Ediacaran Period herald the Cambrian explosion. Geology 47:380384.CrossRefGoogle Scholar
Chen, X., Ling, H. F., Vance, D., Shields-Zhou, G. A., Zhu, M., Poulton, S. W., Och, L. M., et al. 2015. Rise to modern levels of ocean oxygenation coincided with the Cambrian radiation of animals. Nature Communications 6:7142.10.1038/ncomms8142CrossRefGoogle ScholarPubMed
Cherry, L. B., Gilleaudeau, G. J., Grazhdankin, D. V., Romaniello, S. J., Martin, A. J., and Kaufman, A. J.. 2022. A diverse Ediacara assemblage survived under low-oxygen conditions. Nature Communications 13:7306.10.1038/s41467-022-35012-yCrossRefGoogle ScholarPubMed
Cohen, K. M., Finney, S. C., Gibbard, P. L., and Fan, J.-X.. (2013; updated v2022/10). The ICS International Chronostratigraphic Chart. Episodes 36:199204.10.18814/epiiugs/2013/v36i3/002CrossRefGoogle Scholar
Cole, D. B., Mills, D. B., Erwin, D. H., Sperling, E. A., Porter, S. M., Reinhard, C. T., and Planavsky, N. J.. 2020. On the co-evolution of surface oxygen levels and animals. Geobiology 18:260281.10.1111/gbi.12382CrossRefGoogle ScholarPubMed
Crampton, J. S., Beu, A. G., Cooper, R. A., Jones, C. M., Marshall, B., and Maxwell, P. A.. 2003. Estimating the rock volume bias in paleobiodiversity studies. Science 301:358360.10.1126/science.1085075CrossRefGoogle ScholarPubMed
Cunningham, J. A., Vargas, K., Yin, Z., Bengtson, S., and Donoghue, P. C. J.. 2017. The Weng'an Biota (Doushantuo Formation): an Ediacaran window on soft-bodied and multicellular microorganisms. Journal of the Geological Society of London 174:793802.10.1144/jgs2016-142CrossRefGoogle Scholar
Cuthill, J. F. H. 2022. Ediacaran survivors in the Cambrian: suspicions, denials and a smoking gun. Geological Magazine 159:12101219.10.1017/S0016756821001333CrossRefGoogle Scholar
Darroch, S. A., Smith, E. F., Laflamme, M., and Erwin, D. H.. 2018. Ediacaran extinction and Cambrian explosion. Trends in Ecology and Evolution 33:653663.10.1016/j.tree.2018.06.003CrossRefGoogle ScholarPubMed
Darroch, S. A., Cribb, A. T., Buatois, L. A., Germs, G. J., Kenchington, C. G., Smith, E. F., Mocke, H., et al. 2021. The trace fossil record of the Nama Group, Namibia: exploring the terminal Ediacaran roots of the Cambrian explosion. Earth-Science Reviews 212:103435.CrossRefGoogle Scholar
Droser, M. L., and Gehling, J. G.. 2015. The advent of animals: the view from the Ediacaran. Proceedings of the National Academy of Sciences USA 112:48654870.10.1073/pnas.1403669112CrossRefGoogle ScholarPubMed
Dunhill, A. M., Hannisdal, B., and Benton, M. J.. 2014. Disentangling rock record bias and common-cause from redundancy in the British fossil record. Nature Communications 5:19.CrossRefGoogle ScholarPubMed
Dunn, F. S., Liu, A. G., Grazhdankin, D. V., Vixseboxse, P., Flannery-Sutherland, J., Green, E., Harris, S., Wilby, P. R., and Donoghue, P. C.. 2021. The developmental biology of Charnia and the eumetazoan affinity of the Ediacaran rangeomorphs. Science Advances 7:eabe0291.CrossRefGoogle ScholarPubMed
Ebisuzaki, W. 1997. A method to estimate the statistical significance of a correlation when the data are serially correlated. Journal of Climate 10:21472153.10.1175/1520-0442(1997)010<2147:AMTETS>2.0.CO;22.0.CO;2>CrossRefGoogle Scholar
Erwin, D. H. 2008. Macroevolution of ecosystem engineering, niche construction and diversity. Trends in Ecology and Evolution 23:304310.10.1016/j.tree.2008.01.013CrossRefGoogle ScholarPubMed
Erwin, D. H. 2021. Developmental capacity and the early evolution of animals. Journal of the Geological Society 178:jgs2020245.10.1144/jgs2020-245CrossRefGoogle Scholar
Evans, S. D., Hughes, I. V., Gehling, J. G., and Droser, M. L.. 2020. Discovery of the oldest bilaterian from the Ediacaran of South Australia. Proceedings of the National Academy of Sciences USA 117:78457850.CrossRefGoogle ScholarPubMed
Evans, S. D., Tu, C., Rizzo, A., Surprenant, R. L., Boan, P. C., McCandless, H., Marshall, N., Xiao, S., and Droser, M. L.. 2022. Environmental drivers of the first major animal extinction across the Ediacaran White Sea-Nama transition. Proceedings of the National Academy of Sciences USA 119:e2207475119.10.1073/pnas.2207475119CrossRefGoogle ScholarPubMed
Flowers, R. M., Macdonald, F. A., Siddoway, C. S., and Havranek, R.. 2020. Diachronous development of great unconformities before Neoproterozoic Snowball Earth. Proceedings of the National Academy of Sciences USA 117:1017210180.10.1073/pnas.1913131117CrossRefGoogle ScholarPubMed
Gehling, J. G., and Droser, M. L.. 2018. Ediacaran scavenging as a prelude to predation. Emerging Topics in Life Sciences 2:213222.Google ScholarPubMed
Gehling, J. G., García-Bellido, D. C., Droser, M. L., Tarhan, M. L., and Runnegar, B.. 2019. The Ediacaran–Cambrian transition: sedimentary facies versus extinction. Estudios Geológicos 75:e099.10.3989/egeol.43601.554CrossRefGoogle Scholar
Gilbert, P. U. P. A., Bergmann, K. D., Boekelheide, N., Tambutté, S., Mass, T., Marin, F., Adkins, J. F., et al. 2022. Biomineralization: integrating mechanism and evolutionary history. Science Advances 8:eabl9653.10.1126/sciadv.abl9653CrossRefGoogle ScholarPubMed
Glaessner, M. F. 1959. The oldest fossil faunas of South Australia. Geologische Rundschau 47:522531.10.1007/BF01800671CrossRefGoogle Scholar
Hannisdal, B., and Peters, S. E.. 2011. Phanerozoic Earth system evolution and marine biodiversity. Science 334:11211124.CrossRefGoogle ScholarPubMed
Haq, B. U., and Schutter, S. R.. 2008. A chronology of Paleozoic sea-level changes. Science 322:6468.10.1126/science.1161648CrossRefGoogle ScholarPubMed
Haq, B. U., Hardenbol, J. A. N., and Vail, P. R.. 1987. Chronology of fluctuating sea levels since the Triassic. Science 235:11561167.10.1126/science.235.4793.1156CrossRefGoogle ScholarPubMed
He, T., Zhu, M., Mills, B. J., Wynn, P. M., Zhuravlev, A. Y., Tostevin, R., Pogge von Strandmann, P. A., Yang, A., Poulton, S. W., and Shields, G. A.. 2019. Possible links between extreme oxygen perturbations and the Cambrian radiation of animals. Nature Geoscience 12:468474.CrossRefGoogle ScholarPubMed
Heim, N. A., and Peters, S. E.. 2011. Covariation in macrostratigraphic and macroevolutionary patterns in the marine record of North America. Geological Society of America Bulletin 123:620630.10.1130/B30215.1CrossRefGoogle Scholar
Holland, S. M. 2017. Structure, not bias. Journal of Paleontology 91:13151317.CrossRefGoogle Scholar
Husson, J. M., and Peters, S.E.. 2018. Nature of the sedimentary rock record and its implications for Earth system evolution. Emerging Topics in Life Sciences 2:125136.Google ScholarPubMed
Jiang, L., Zhao, M., Shen, A., Huang, L., Chen, D., and Cai, C.. 2022. Pulses of atmosphere oxygenation during the Cambrian radiation of animals. Earth and Planetary Science Letters 590:117565.10.1016/j.epsl.2022.117565CrossRefGoogle Scholar
Keller, C. B., Husson, J. M., Mitchell, R. N., Bottke, W. F., Gernon, T. M., Boehnke, P., Bell, E. A., Swanson-Hysell, N. L., and Peters, S. E.. 2019. Neoproterozoic glacial origin of the Great Unconformity. Proceedings of the National Academy of Sciences USA 116:1361145.CrossRefGoogle ScholarPubMed
Knoll, A. H. 2003. Biomineralization and evolutionary history. Reviews in Mineralogy and Geochemistry 54:329356.10.2113/0540329CrossRefGoogle Scholar
Knoll, A. H., and Carroll, S. B.. 1999. Early animal evolution: emerging views from comparative biology and geology. Science 284:21292137.10.1126/science.284.5423.2129CrossRefGoogle Scholar
Knoll, A. H., Walter, M. R., Narbonne, G. M., and Christie-Blick, N.. 2006. The Ediacaran Period: a new addition to the geologic time scale. Lethaia 39:1330.10.1080/00241160500409223CrossRefGoogle Scholar
Laflamme, M., Darroch, S. A., Tweedt, S. M., Peterson, K. J., and Erwin, D. H.. 2013. The end of the Ediacara biota: extinction, biotic replacement, or Cheshire Cat? Gondwana Research 23:558573.CrossRefGoogle Scholar
Lenton, T. M., Boyle, R. A., Poulton, S. W., Shields-Zhou, G. A., and Butterfield, N. J.. 2014. Co-evolution of eukaryotes and ocean oxygenation in the Neoproterozoic era. Nature Geoscience 7:257265.10.1038/ngeo2108CrossRefGoogle Scholar
Li, C., Shi, W., Cheng, M., Jin, C., and Algeo, T. J.. 2020. The redox structure of Ediacaran and early Cambrian oceans and its controls. Science Bulletin 65:21412149.10.1016/j.scib.2020.09.023CrossRefGoogle Scholar
Liu, A. G., and Tindal, B. H.. 2021. Ediacaran macrofossils prior to the ~580 Ma Gaskiers glaciation in Newfoundland, Canada. Lethaia 54:260270.10.1111/let.12401CrossRefGoogle Scholar
Lorentzen, S., Augustsson, C., Nystuen, J. P., Berndt, J., Jahren, J., and Schovsbo, N. H.. 2018. Provenance and sedimentary processes controlling the formation of lower Cambrian quartz arenite along the southwestern margin of Baltica. Sedimentary Geology 375:203217.10.1016/j.sedgeo.2017.08.008CrossRefGoogle Scholar
Macdonald, F. A., Yonkee, W. A., Flowers, R. M., and Swanson-Hysell, N. L.. 2023. Neoproterozoic of Laurentia. In Whitmeyer, S. J., Williams, M. L., Kellett, D. A., and Tikoff, B., eds. Turning points in the evolution of a continent. Geological Society of America Bulletin 220:9780813782201. Geological Society of America, Boulder, Colo.Google Scholar
Mackenzie, F. T., and Pigott, J. D.. 1981. Tectonic controls of Phanerozoic sedimentary rock cycling. Journal of the Geological Society of London 138:183196.10.1144/gsjgs.138.2.0183CrossRefGoogle Scholar
McDannell, K. T., and Keller, C. B.. 2022. Cryogenian glacial erosion of the central Canadian Shield: the “late” Great Unconformity on thin ice. Geology 50:13361340.CrossRefGoogle Scholar
McGowan, A. J., and Smith, A. B.. 2008. Are global Phanerozoic marine diversity curves truly global? A study of the relationship between regional rock records and global Phanerozoic marine diversity. Paleobiology 34:80103.10.1666/07019.1CrossRefGoogle Scholar
Meyers, S. R. 2014. Astrochron: an R package for astrochronology. https://cran.r-project.org/package=astrochron, accessed 11 March 2023.Google Scholar
Meyers, S. R., and Peters, S. E.. 2011. A 56 million year rhythm in North American sedimentation during the Phanerozoic. Earth and Planetary Science Letters 303:174180.10.1016/j.epsl.2010.12.044CrossRefGoogle Scholar
Miall, A. D. 2016. The valuation of unconformities. Earth-Science Reviews 163:2271.CrossRefGoogle Scholar
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., and Pekar, S. F.. 2005. The Phanerozoic record of global sea-level change. Science 310:12931298.10.1126/science.1116412CrossRefGoogle Scholar
Muscente, A. D., Bykova, N., Boag, T. H., Buatois, L. A., Mángano, M. G., Eleish, A., Prabhu, A., et al. 2019. Ediacaran biozones identified with network analysis provide evidence for pulsed extinctions of early complex life. Nature Communications 10:115.10.1038/s41467-019-08837-3CrossRefGoogle ScholarPubMed
Na, L., Kocsis, Á. T., Li, Q., and Kiessling, W.. 2022. Coupling of geographic range and provincialism in Cambrian marine invertebrates. Paleobiology 49:284295.CrossRefGoogle Scholar
Nance, R. D., Murphy, J. B., and Santosh, M.. 2014. The supercontinent cycle: a retrospective essay. Gondwana Research 25:429.10.1016/j.gr.2012.12.026CrossRefGoogle Scholar
Nawrot, R., Scarponi, D., Azzarone, M., Dexter, T. A., Kusnerik, K. M., Wittmer, J. M., Amorosi, A., and Kowalewski, M.. 2018. Stratigraphic signatures of mass extinctions: Ecological and sedimentary determinants. Proceedings of the Royal Society of London B 285:20181191.Google ScholarPubMed
Newell, N. D. 1959. The nature of the fossil record. Proceedings of the American Philosophical Society 103:264285.Google Scholar
Och, L. M., and Shields-Zhou, G. A.. 2012. The Neoproterozoic oxygenation event: environmental perturbations and biogeochemical cycling. Earth-Science Reviews 110:2657.10.1016/j.earscirev.2011.09.004CrossRefGoogle Scholar
Oksanen, J., Simpson, G. L., Blanchet, F. G., Kindt, R., Legendre, P. R., Minchin, P., O'Hara, R. B., et al. 2022. vegan: community ecology package, R package version 2.6-4. https://CRAN.R-project.org/package=vegan, accessed 11 March 2023.Google Scholar
Peak, B. A., Flowers, R. M., and Macdonald, F. A.. 2023. Ediacaran–Ordovician tectonic and geodynamic drivers of Great Unconformity exhumation on the southern Canadian Shield. Earth and Planetary Science Letters 619:118334.CrossRefGoogle Scholar
Peters, S. E. 2005. Geologic constraints on the macroevolutionary history of marine animals. Proceedings of the National Academy of Sciences USA 102:1232612331.CrossRefGoogle ScholarPubMed
Peters, S. E. 2006. Genus extinction, origination, and the durations of sedimentary hiatuses. Paleobiology 32:387407.10.1666/05081.1CrossRefGoogle Scholar
Peters, S. E. 2008. Environmental determinants of extinction selectivity in the fossil record. Nature 454:626629.10.1038/nature07032CrossRefGoogle Scholar
Peters, S. E., and Foote, M.. 2001. Biodiversity in the Phanerozoic: a reinterpretation. Paleobiology 27:583601.10.1666/0094-8373(2001)027<0583:BITPAR>2.0.CO;22.0.CO;2>CrossRefGoogle Scholar
Peters, S. E., and Foote, M.. 2002. Determinants of extinction in the fossil record. Nature 416:420424.10.1038/416420aCrossRefGoogle ScholarPubMed
Peters, S. E., and Gaines, R. R.. 2012. Formation of the “Great Unconformity” as a trigger for the Cambrian Explosion. Nature 484:363366.10.1038/nature10969CrossRefGoogle ScholarPubMed
Peters, S. E., and Heim, N. A.. 2011a. Macrostratigraphy and macroevolution in marine environments: testing the common-cause hypothesis In McGowan, A. J. and Smith, A. B., eds. Comparing the geological and fossil records: implications for biodiversity studies. Geological Society of London Special Publication 358:95104. Geological Society of London, London.Google Scholar
Peters, S. E., and Heim, N. A.. 2011b. Stratigraphic distribution of marine fossils in North America. Geology 39:259262.10.1130/G31442.1CrossRefGoogle Scholar
Peters, S. E., and Husson, J. M.. 2017. Sediment cycling on continental and oceanic crust. Geology 45:323326.10.1130/G38861.1CrossRefGoogle Scholar
Peters, S. E., and McClennen, M.. 2016. The Paleobiology Database application programming interface. Paleobiology 42:17.10.1017/pab.2015.39CrossRefGoogle Scholar
Peters, S. E., Kelly, D. C., and Fraass, A. J.. 2013. Oceanographic controls on the diversity and extinction of planktonic foraminifera. Nature 493:398401.10.1038/nature11815CrossRefGoogle ScholarPubMed
Peters, S. E., Husson, J. M., and Czaplewski, J.. 2018. Macrostrat: a platform for geological data integration and deep-time Earth crust research. Geochemistry, Geophysics, Geosystems 19:13931409.10.1029/2018GC007467CrossRefGoogle Scholar
Peters, S. E., Quinn, D. P., Husson, J. M., and Gaines, R. R.. 2022. Macrostratigraphy: insights into cyclic and secular evolution of the Earth-life system. Annual Review of Earth and Planetary Sciences 50:419449.10.1146/annurev-earth-032320-081427CrossRefGoogle Scholar
Pu, J. P., Bowring, S. A., Ramezani, J., Myrow, P., Raub, T. D., Landing, E., Mills, A., Hodgin, E., and Macdonald, F. A.. 2016. Dodging snowballs: geochronology of the Gaskiers glaciation and the first appearance of the Ediacaran biota. Geology 44:955958.10.1130/G38284.1CrossRefGoogle Scholar
Raup, D. M. 1972. Taxonomic diversity during the Phanerozoic. Science 177:10651071.10.1126/science.177.4054.1065CrossRefGoogle ScholarPubMed
Raup, D. M. 1976. Species diversity in the Phanerozoic: an interpretation. Paleobiology 2:289297.10.1017/S0094837300004929CrossRefGoogle Scholar
Raup, D.M., and Sepkoski, J. J. Jr. 1982. Mass extinctions in the marine fossil record. Science 215:15011503.10.1126/science.215.4539.1501CrossRefGoogle ScholarPubMed
Rook, D. L., Heim, N. A., and Marcot, J.. 2013. Contrasting patterns and connections of rock and biotic diversity in the marine and non-marine fossil records of North America. Palaeogeography, Palaeoclimatology, Palaeoecology 372:123129.10.1016/j.palaeo.2012.10.006CrossRefGoogle Scholar
Rooney, A. D., Cantine, M. D., Bergmann, K. D., Gómez-Pérez, I., Al Baloushi, B., Boag, T. H., Busch, J. F., Sperling, E. A., and Strauss, J. V.. 2020. Calibrating the coevolution of Ediacaran life and environment. Proceedings of the National Academy of Sciences USA 117:1682416830.10.1073/pnas.2002918117CrossRefGoogle ScholarPubMed
Schiffbauer, J. D., Huntley, J. W., O'Neil, G. R., Darroch, S. A., Laflamme, M., and Cai, Y.. 2016. The latest Ediacaran wormworld fauna: setting the ecological stage for the Cambrian explosion. GSA Today 26:411.10.1130/GSATG265A.1CrossRefGoogle Scholar
Sears, J. W., and Price, R. A.. 2003. Tightening the Siberian connection to western Laurentia. Geological Society of America Bulletin 115:943953.10.1130/B25229.1CrossRefGoogle Scholar
Segessenman, D. C., and Peters, S. E.. 2023. Macrostratigraphy of the Ediacaran System in North America, Laurentia. In Whitmeyer, S. J., Williams, M. L., Kellett, D. A., and Tikoff, B., eds. Turning points in the evolution of a continent. Geological Society of America Bulletin 220:9780813782201. Geological Society of America, Boulder, Colo.Google Scholar
Seilacher, R. W. 1984. Late Precambrian and early Cambrian metazoa: preservational or real extinctions? Pp. 159168 in Holland, H. D. and Trendall, A. F., eds. Patterns of change in Earth evolution. Springer, Berlin.10.1007/978-3-642-69317-5_10CrossRefGoogle Scholar
Sepkoski, J. J., Bambach, R. K., Raup, D. M., and Valentine, J. W.. 1981. Phanerozoic marine diversity and the fossil record. Nature 293:435437.10.1038/293435a0CrossRefGoogle Scholar
Servais, T., Cascales-Miñana, B., Harper, D. A. T., Lefebvre, B., Munnecke, A., Wang, W., and Zhang, Y.. 2023. No (Cambrian) explosion and no (Ordovician) event: a single long-term radiation in the early Palaeozoic. Palaeogeography, Palaeclimatology, Palaeocology 623:111592.10.1016/j.palaeo.2023.111592CrossRefGoogle Scholar
Shahkarami, S., Buatois, L. A., Mángano, M. G., Hagadorn, J. W., and Almond, J.. 2020. The Ediacaran–Cambrian boundary: evaluating stratigraphic completeness and the Great Unconformity. Precambrian Research 345:105721.10.1016/j.precamres.2020.105721CrossRefGoogle Scholar
Shannon, C. E. 1948. A mathematical theory of communication. Bell System Technical Journal 27:379423.10.1002/j.1538-7305.1948.tb01338.xCrossRefGoogle Scholar
Shore, A. J., Wood, R. A., Butler, I. B., Zhuravlev, A. Y., McMahon, S., Curtis, A., and Bowyer, F. T.. 2021. Ediacaran metazoan reveals lophotrochozoan affinity and deepens root of Cambrian Explosion. Science Advances 7:eabf2933.CrossRefGoogle ScholarPubMed
Sloss, L. L. 1963. Sequences in the cratonic interior of North America. Geological Society of America Bulletin 74:93114.10.1130/0016-7606(1963)74[93:SITCIO]2.0.CO;2CrossRefGoogle Scholar
Smith, A. B. 2001. Large–scale heterogeneity of the fossil record: implications for Phanerozoic biodiversity studies. Philosophical Transactions of the Royal Society of London B 356:351367.10.1098/rstb.2000.0768CrossRefGoogle ScholarPubMed
Smith, A. B. 2007. Marine diversity through the Phanerozoic: problems and prospects. Journal of the Geological Society of London 164:731745.10.1144/0016/76492006-184CrossRefGoogle Scholar
Smith, A. B., and McGowan, A. J.. 2007. The shape of the Phanerozoic marine palaeodiversity curve: how much can be predicted from the sedimentary rock record of Western Europe? Palaeontology 50:765774.10.1111/j.1475-4983.2007.00693.xCrossRefGoogle Scholar
Smith, A. B., Gale, A. S., and Monks, N. E. A.. 2001. Sea-level change and rock-record bias in the Cretaceous: a problem for extinction and biodiversity studies. Paleobiology 27:241253.10.1666/0094-8373(2001)027<0241:SLCARR>2.0.CO;22.0.CO;2>CrossRefGoogle Scholar
Sperling, E. A., Carbone, C., Strauss, J. V., Johnston, D. T., Narbonne, G. M., and Macdonald, F. A.. 2016. Oxygen, facies, and secular controls on the appearance of Cryogenian and Ediacaran body and trace fossils in the Mackenzie Mountains of northwestern Canada. Geological Society of America Bulletin 128:558575.10.1130/B31329.1CrossRefGoogle Scholar
Sprigg, R. C. 1947. Early Cambrian (?) jellyfishes from the Flinders Ranges, South Australia. Transactions of the Royal Society of South Australia 71:212224.Google Scholar
Stanley, S. M. 2007. An analysis of the history of marine animal diversity. Paleobiology 33(S4):155.Google Scholar
Sturrock, C. P., Flowers, R. M., and Macdonald, F. A.. 2021. The late Great Unconformity of the Central Canadian Shield. Geochemistry, Geophysics, Geosystems 22:e2020GC009567.10.1029/2020GC009567CrossRefGoogle Scholar
Tarhan, L. G., Droser, M. L., Cole, D. B., and Gehling, J. G.. 2018. Ecological expansion and extinction in the late Ediacaran: weighing the evidence for environmental and biotic drivers. Integrative and Comparative Biology 58:688702.10.1093/icb/icy020CrossRefGoogle Scholar
Tarhan, L. G., Myrow, P. M., Smith, E. F., Nelson, L. L., and Sadler, P. M.. 2020. Infaunal augurs of the Cambrian explosion: an Ediacaran trace fossil assemblage from Nevada, USA. Geobiology 18:486496.10.1111/gbi.12387CrossRefGoogle ScholarPubMed
Tasistro-Hart, A. R., and Macdonald, F. A.. 2023. Phanerozoic flooding of North America and the Great Unconformity. Proceedings of the National Academy of Sciences USA 120:e2309084120.CrossRefGoogle ScholarPubMed
Valentine, J. W. 1969. Patterns of taxonomic and ecological structure of the shelf benthos during Phanerozoic time. Palaeontology 12:684709.Google Scholar
Waggoner, B. 2003. The Ediacaran biotas in space and time. Integrative and Comparative Biology 43:104113.10.1093/icb/43.1.104CrossRefGoogle ScholarPubMed
Wang, D., Ling, H. F., Struck, U., Zhu, X. K., Zhu, M., He, T., Yang, B., Gamper, A., and Shields, G. A.. 2018. Coupling of ocean redox and animal evolution during the Ediacaran–Cambrian transition. Nature Communications 9:2575.10.1038/s41467-018-04980-5CrossRefGoogle ScholarPubMed
Weldeghebriel, M. F., Lowenstein, T. K., García-Veigas, J., and Cendón, D.I.. 2022. [Ca2+] and [SO42−] in Phanerozoic and terminal Proterozoic seawater from fluid inclusions in halite: the significance of Ca-SO4 crossover points. Earth and Planetary Science Letters 594:117712.10.1016/j.epsl.2022.117712CrossRefGoogle Scholar
Wood, R., Liu, A. G., Bowyer, F., Wilby, P. R., Dunn, F. S., Kenchington, C. G., Cuthill, J. F. H., Mitchell, E. G., and Penny, A.. 2019. Integrated records of environmental change and evolution challenge the Cambrian Explosion. Nature Ecology and Evolution 3:528538.10.1038/s41559-019-0821-6CrossRefGoogle ScholarPubMed
Xiao, S. H., and Laflamme, M.. 2009. On the eve of animal radiation: phylogeny, ecology and evolution of the Ediacara biota. Trends in Ecology and Evolution 24:3140.10.1016/j.tree.2008.07.015CrossRefGoogle ScholarPubMed
Xiao, S. H., and Narbonne, G. M.. 2020. The Ediacaran Period. Pp. 521561 in Gradstein, F. M., Ogg, J. G., Schmitz, M. D., and Ogg, G. M., eds. Geologic time scale 2020. Elsevier, Cambridge, Mass.10.1016/B978-0-12-824360-2.00018-8CrossRefGoogle Scholar
Yang, C., Rooney, A. D., Condon, D. J., Li, X. H., Grazhdankin, D. V., Bowyer, F. T., Hu, C., Macdonald, F. A., and Zhu, M.. 2021. The tempo of Ediacaran evolution. Science Advances 7:eabi9643.10.1126/sciadv.abi9643CrossRefGoogle ScholarPubMed
Yang, C., Li, Y., Selby, D., Wan, B., Guan, C., Zhou, C., and Li, X. H.. 2022. Implications for Ediacaran biological evolution from the ca. 602 Ma Lantian biota in China. Geology 50:562566.Google Scholar
Zhang, F., Xiao, S. H., Romaniello, S. J., Hardisty, D., Li, C., Melezhik, V., Pokrovsky, B., et al. 2019. Global marine redox changes drove the rise and fall of the Ediacara biota. Geobiology 17:594610.10.1111/gbi.12359CrossRefGoogle ScholarPubMed
Zhang, G., Chen, D., Huang, K. J., Liu, M., Huang, T., Yeasmin, R., and Fu, Y.. 2021. Dramatic attenuation of continental weathering during the Ediacaran–Cambrian transition: implications for the climatic-oceanic-biological co-evolution. Global and Planetary Change 203:103518.CrossRefGoogle Scholar
Zhang, X., and Shu, D.. 2021. Current understanding on the Cambrian Explosion: questions and answers. PalZ 95:641660.10.1007/s12542-021-00568-5CrossRefGoogle Scholar
Zhuravlev, A. Y., and Wood, R. A.. 2018. The two phases of the Cambrian Explosion. Scientific Reports 8:16656.10.1038/s41598-018-34962-yCrossRefGoogle ScholarPubMed
Zhuravlev, A. Y., and Wood, R.. 2020. Dynamic and synchronous changes in metazoan body size during the Cambrian Explosion. Scientific Reports 10:18.10.1038/s41598-020-63774-2CrossRefGoogle ScholarPubMed
Zou, Y., Mitchell, R. N., Chu, X., Brown, M., Jiang, J., Li, Q., Zhao, L., and Zhai, M.. 2023. Surface evolution during the mid-Proterozoic stalled by mantle warming under Columbia–Rodinia. Earth and Planetary Science Letters 607:118055.10.1016/j.epsl.2023.118055CrossRefGoogle Scholar
Figure 0

Figure 1. Maps of North America with sediment-bearing column areas from Macrostrat (colored polygons) and fossil collection locations from the Paleobiology Database (PBDB; blue diamonds). Fossil collection locations have been randomly offset by a factor of 0.5%. Total numbers of columns and fossil collections are shown on each map. A, Lower Ediacaran (635–590 Ma); B, upper Ediacaran (590–538.8 Ma); C, Terreneuvian (538.8–521 Ma); D, Series 2 (521–509 Ma); E, Miaolingian (509–497 Ma); F, Furongian (497–485.4 Ma). Lower Ediacaran and upper Ediacaran informal divisions based on initial rise in preserved sediment area and volume. Cambrian epoch timings based on Cohen et al. (2013; updated v2022/10).

Figure 1

Figure 2. Time series of rock and fossil metrics from the Ediacaran and Cambrian with Mackenzie and Sauk Sloss sequences and the Ediacaran/Cambrian transition highlighted. A, Log-scale plot of the number of occurrences and genera with bootstrap resampling–generated confidence intervals (2σ); B, log-scale plot of the total number of sedimentary rock units and the number of rock units that contain at least one fossil occurrence with bootstrap resampling–generated confidence intervals (2σ); C, stacked area plot of preserved marine sedimentary rock area (km2) divided into clastic (grain size–based) and carbonate categories; D, stacked area plot of calculated marine sedimentary volume flux (km3/Myr); and E, proportion of fossil-occupied sedimentary units (black), area (blue), and volume flux (red). Note that pre–580 Ma occurrences include fossil data from thicker, undivided stratigraphic sections with few geochronologic constraints.

Figure 2

Figure 3. Time series of median thickness and duration of sedimentary rocks and the number of occurrences reported from clastic and carbonate lithologies with Mackenzie and Sauk Sloss sequences and the Ediacaran/Cambrian transition highlighted. A, Median thickness (m) of all sedimentary units (black) with bootstrap resampling–generated confidence interval (2σ) and only sedimentary units that are occupied (contain at least one occurrence; red); B, median duration (Myr) of all sedimentary units (black) with bootstrap resampling–generated confidence interval (2σ) and only sedimentary units that are occupied (red); and C, stacked area of occurrence counts by Paleobiology Database (PBDB) reported lithology. A single occurrence can have multiple lithologies and therefore can be counted within multiple lithologic categories for one time interval.

Figure 3

Figure 4. Time series of occurrence and genus counts normalized by sedimentary rock units (occurrences/rock unit), rock area (occurrences/10,000 km2), and volume flux (occurrences/1000 km3/Myr) with Mackenzie and Sauk Sloss sequences and the Ediacaran/Cambrian transition highlighted. A, Counts of occurrences normalized to sedimentary rock quantities; and B, counts of genera normalized to sedimentary rock quantities. See text for discussion.

Figure 4

Figure 5. Time series of raw genus counts and the Shannon H indices of unique genus names and their reported locations (“states” field in the Paleobiology Database [PBDB]) with Mackenzie and Sauk Sloss sequences and the Ediacaran/Cambrian transition highlighted. A, Raw genus counts; B, Shannon H index of unique genera names and their occurrence frequencies; and C, Shannon H index of unique state names and their occurrence frequencies. Decreases in Shannon H indices from ca. 565–555 Ma and ca. 515–505 Ma represent intervals in which sampling is dominated by collections at specific localities (Mistaken Point Fm., Newfoundland, Canada, and Stephen Fm., British Columbia, Canada, respectively).

Figure 5

Table 1. Spearman rank correlation coefficients, ρ (rs), and associated p-values calculated between Ediacaran–Cambrian sedimentary rock and fossil quantities in 1, 5, and 10 Myr bins. Green cells represent correlation coefficients with corresponding p-values in the 95% confidence interval (p < 0.05), and yellow cells represent the 90% confidence interval (p < 0.1). Bold values are Spearman's rho (ρ) rank correlation coefficients. “No. of sed. marine units” is the number of sedimentary marine units.