A ‘Neoproterozoic oxygenation event’ is widely invoked as a causal factor in animal evolution, and often attributed to abiotic causes such as post-glacial pulses of phosphorus weathering. However, recent evidence suggests a series of transient ocean oxygenation events ∼660–520 Ma, which do not fit the simple model of a monotonic rise in atmospheric oxygen (pO2). Hence, we consider mechanisms by which the evolution of marine eukaryotes, coupled with biogeochemical and ecological feedbacks, potentially between alternate stable states, could have caused changes in ocean carbon cycling and redox state, phosphorus cycling and atmospheric pO2. We argue that the late Tonian ocean ∼750 Ma was dominated by rapid microbial cycling of dissolved organic matter (DOM) with elevated nutrient (P) levels due to inefficient removal of organic matter to sediments. We suggest the abrupt onset of the eukaryotic algal biomarker record ∼660–640 Ma was linked to an escalation of protozoan predation, which created a ‘biological pump’ of sinking particulate organic matter (POM). The resultant transfer of organic carbon (Corg) and phosphorus to sediments was strengthened by subsequent eukaryotic innovations, including the advent of sessile benthic animals and mobile burrowing animals. Thus, each phase of eukaryote evolution tended to lower P levels and oxygenate the ocean on ∼104 year timescales, but by decreasing Corg/P burial ratios, tended to lower atmospheric pO2 and deoxygenate the ocean again on ∼106 year timescales. This can help explain the transient nature and ∼106 year duration of oceanic oxygenation events through the Cryogenian–Ediacaran–Cambrian.

Marine eukaryotes are important ecosystem engineers and their evolution over ∼850–500 Ma (Figure 1a,b) surely had an impact on carbon, phosphorus and oxygen cycling [1]. While a monotonic rise in atmospheric oxygen (pO2) in a ‘Neoproterozoic oxygenation event’ [25] is still regularly invoked to explain evidence of deeper ocean oxygenation in the Neoproterozoic Era, ocean oxygenation could equally have been caused by declining oxygen demand in deeper waters [1,6]. Here, ‘deeper’ could mean anything below the well-mixed surface layer of stratified shelf seas or the open ocean, but sampling of deep time is largely restricted to continental shelf-slope ocean margins (so does not extend to the deep open ocean) [6]. In deeper waters, oxygenation state is governed by the balance of oxygen supply and respiratory demand from organic matter input. Declining oxygen demand could occur due to: (i) a redistribution of organic matter (and its remineralisation) away from these depths, and/or (ii) a global decline in ocean nutrient levels and organic matter primary production. We focus on phosphorus as the ultimate limiting nutrient, because the ocean available nitrogen inventory — whether as , or a mix of both [6] — tends to track changes in the phosphorus inventory [7], even though models predict nitrogen would have been much further below the ‘Redfield ratio’ to phosphorus (required by organisms) in a more anoxic ocean [6,8,9].

Timeline of biological and environmental changes through the Tonian, Cryogenian, Ediacaran and Cambrian periods (850–500 Ma).

Figure 1.
Timeline of biological and environmental changes through the Tonian, Cryogenian, Ediacaran and Cambrian periods (850–500 Ma).

(a) Biomarker evidence [34]; sterane/hopane ratio indicates balance of eukaryotes/bacteria, followed by occurrence of eukaryotic steranes (with probable sources [34,35]) including cholestane (heterotrophic eukaryotes), stigmastane (chlorophytes), ergostane (general), cryostane (unknown), 24-ipc = 24-isopropylcholestane (demosponges), 24-npc = 24-n-propylcholestane (rhizarians). (b) Fossil evidence, updated from [1] based on studies cited in main text and [33]. (c) Ocean redox state from redox-sensitive elements (pink; global signature) [11,13] and iron speciation data at different depths (green, yellow, orange; local signature) [13,109,110]. (d) The carbon isotopic composition of marine carbonates.

Figure 1.
Timeline of biological and environmental changes through the Tonian, Cryogenian, Ediacaran and Cambrian periods (850–500 Ma).

(a) Biomarker evidence [34]; sterane/hopane ratio indicates balance of eukaryotes/bacteria, followed by occurrence of eukaryotic steranes (with probable sources [34,35]) including cholestane (heterotrophic eukaryotes), stigmastane (chlorophytes), ergostane (general), cryostane (unknown), 24-ipc = 24-isopropylcholestane (demosponges), 24-npc = 24-n-propylcholestane (rhizarians). (b) Fossil evidence, updated from [1] based on studies cited in main text and [33]. (c) Ocean redox state from redox-sensitive elements (pink; global signature) [11,13] and iron speciation data at different depths (green, yellow, orange; local signature) [13,109,110]. (d) The carbon isotopic composition of marine carbonates.

Close modal

Available data show that ocean redox conditions were temporally (as well as spatially) variable in the late Neoproterozoic-early Paleozoic (Figure 1c), while the carbon cycle underwent major fluctuations (Figure 1d). A series of transient ocean oxygenation events are inferred, beginning in the middle of the Cryogenian period (720–635 Ma) [10,11], and getting more frequent through the Ediacaran (635–540 Ma) [12] and early-mid Cambrian (540–500 Ma) [13]. If these transient events were due purely to changes in atmospheric pO2, then it must have peaked and declined repeatedly. They could equally have been caused by rapid reorganizations of nutrient and carbon cycling within the ocean, potentially between alternative stable states (anoxic high P recycling, oxic efficient P removal) [14]. Indeed, the onset of oxygenation appears sufficiently rapid in some cases [12] that ∼104 year declines in ocean nutrient inventories look a more plausible explanation than ∼106 year increases in atmospheric pO2. These are not mutually exclusive scenarios as rapid biogeochemical reorganisations would ultimately be countered by a slower adjustment of pO2 [14].

Here, we explore how eukaryote evolution could have changed the nutrient and redox state of their (submarine) world. We start by introducing key controls on ocean redox state, marine phosphorus cycling and oxygen cycling. Then, we consider possible causes of the rise to ecological prominence of eukaryotic algae, their biogeochemical effects, and those of early sessile animals, and later mobile burrowing animals.

Timescales and processes of eukaryotic effects on biogeochemical cycling.

Figure 2.
Timescales and processes of eukaryotic effects on biogeochemical cycling.

(a) Ocean circulation (∼103 year) timescale (black arrows): circulation redistributes O2 supply from atmosphere and O2 demand from DOM (green dotted background) and sinking POM (green downward wiggle arrow) within the water column. (b) Phosphorus cycling (∼104 year) timescale (dark blue arrows), inset of shelf sea: the ocean P inventory adjusts to maintain balance between P input (via weathering) and P burial (primarily on shelves). P sequestration in sediments is enhanced by biological pump (downward wiggle arrow) and by sessile benthic animals (downward arrow). P is preferentially recycled from sediments especially under euxinic conditions (upward arrow). (c) Oxygen cycling (∼106 year) timescale (red arrows): Organic carbon burial (Corg; green downward arrow) provides O2 source, governed by P input and Corg:P burial ratio which is redox-sensitive. Oxidative weathering of ancient Corg in sedimentary rocks (green upward arrow) provides O2 sink.

Figure 2.
Timescales and processes of eukaryotic effects on biogeochemical cycling.

(a) Ocean circulation (∼103 year) timescale (black arrows): circulation redistributes O2 supply from atmosphere and O2 demand from DOM (green dotted background) and sinking POM (green downward wiggle arrow) within the water column. (b) Phosphorus cycling (∼104 year) timescale (dark blue arrows), inset of shelf sea: the ocean P inventory adjusts to maintain balance between P input (via weathering) and P burial (primarily on shelves). P sequestration in sediments is enhanced by biological pump (downward wiggle arrow) and by sessile benthic animals (downward arrow). P is preferentially recycled from sediments especially under euxinic conditions (upward arrow). (c) Oxygen cycling (∼106 year) timescale (red arrows): Organic carbon burial (Corg; green downward arrow) provides O2 source, governed by P input and Corg:P burial ratio which is redox-sensitive. Oxidative weathering of ancient Corg in sedimentary rocks (green upward arrow) provides O2 sink.

Close modal

We distinguish three timescales over which eukaryotic evolutionary or ecologically driven changes are coupled to biogeochemical cycles (Figure 2):

  • (1) A ∼103 year ocean circulation timescale [15] of effects on the nature and distribution of organic matter and corresponding oxygen demand in the water column (and between it and the sediments; Figure 2a).

  • (2) A ∼104 year timescale [16] of changes in P burial efficiency altering the ocean P inventory, which in turn controls the amount of organic matter produced and corresponding oxygen demand (Figure 2b). This modern P timescale could have differed in the Neoproterozoic.

  • (3) A ∼106 year timescale of changes in the oxygen source from organic carbon (Corg) burial altering atmospheric pO2 (Figure 2c), which in turn controls oxygen supply to the water column. This timescale is estimated for the Neoproterozoic based on pO2 ∼0.1 PAL (present atmospheric level) supported by ∼2.5 × 1012 mol year−1 net O2 source [17].

When looking at the low temporal resolution of the geological record, we expect that the effects of eukaryote evolution on organic matter cycling (1) cannot be resolved from their effects on phosphorus cycling (2), whereas the slower timescale of changes in atmospheric pO2 (3) is resolvable from the others.

Today, a relatively efficient ‘biological pump’ of sinking particulate organic matter (POM) creates oxygen demand across a range of water depths, and transfers Corg and P to sediments.

On geologic timescales >104 years (i.e. longer than the residence time of P in the ocean), global phosphorus output to sediments must equal global input from rivers (ultimately derived from weathering). This means that changes in P input or the efficiency of P removal (the focus here) cause the ocean P inventory and corresponding output flux to adjust until output again matches input. Today, despite an efficient biological pump, ∼70% of global phosphorus removal occurs in shelf sea settings and only ∼30% on the deep ocean floor [16,18]. Hence, here we ignore P burial in deep ocean sediments and focus on shelf seas, because P burial there largely controls the global P reservoir.

The burial of photosynthetically derived Corg also occurs mostly in shelf seas [19] and represents the major net source of oxygen to the atmosphere: CO2+H2O→CH2O(↓buried)+O2(↑). The Corg burial flux (FCorg) is controlled by the phosphorus input/burial flux (FP; which limits production) and the burial ratio of Corg to total phosphorus in marine sediments [(Corg/P)burial; which is redox-sensitive]: FP × (Corg/P)burial = FCorg. To first order, we do not expect evolutionary or ecologically driven changes in organic matter cycling within the ocean to have impacted atmospheric pO2, because the ocean P inventory adjusts to changes in P burial efficiency such that the P burial flux again matches the (unchanged) P input flux — and the corresponding oxygen source from Corg burial is left unchanged.

In reality, the ratio of Corg to total phosphorus buried in new sedimentary rocks, (Corg/P)burial, can change (thus affecting atmospheric pO2), because it is sensitive to the redox state of ocean bottom waters and hence to ocean nutrient levels and (ultimately) atmospheric pO2. Under anoxic and euxinic (sulfate-reducing) bottom-water conditions, phosphorus recycling to the water column is enhanced, increasing (Corg/P)burial [20,21], and oxygenation reduces P recycling from sediments [22] decreasing (Corg/P)burial. This acts as a positive feedback amplifying changes in phosphorus levels, productivity and redox state on ∼104 year timescales [23]. However, on ∼106 year timescales, changes in atmospheric pO2 provide negative feedback [14,24]. Less clear is what happens to phosphorus cycling under widespread anoxic but ‘ferruginous’ (iron-reducing) conditions [25]. While the formation of Fe(III) oxides and associated trapping of P in sediments will be suppressed, there is evidence that vivianite and mixed Fe(II)–Fe(III) minerals (‘green rust’) can provide a potentially large P sink [26,27]. This could change the sign of the feedback, such that ferruginous conditions enhance P removal suppressing their own spread, but amplify declining pO2 [28,29].

Picture the ocean in the mid to late Tonian period (∼850–720 Ma). Despite the evolution of algae ∼1.7–1.4 Ga [30], including multi-cellular red algae by ∼1.1 Ga [31], and the radiation of red and green algae into the marine environment [32,33], they left no biomarker record [34] (Figure 1a). The low sterane/hopane ratio of biomarkers [34] indicates that bacteria dominated preserved lipids, with cholestane (probably from heterotrophic eukaryotes) and cryostane (of unknown origin [35]) the only eukaryotic steranes present. This requires that unicellular algae were either an ecologically insignificant contributor to marine productivity [34], or that eukaryotic biomass was efficiently recycled either in the water column [36] or in microbial mats [37]. Efficient water column recycling of primary production by prokaryotes (Figure 3a) and possibly small protists (Figure 3c) would have closed a ‘microbial loop’ ensuring dissolved organic matter (DOM; operationally defined as <0.22 or <0.7 µm) [38,39] dominated the water column. This DOM pool was not a massive, ancient one [40], rather it turned over on <104 year timescales consistent with modern observations [38,39].

Community size structure in an idealized steady-state microbial food chain model [52,111], for assumed evolutionary steps. Green = autotrophs, brown = protist heterotrophs, blue = dissolved nitrogen.

Figure 3.
Community size structure in an idealized steady-state microbial food chain model [52,111], for assumed evolutionary steps. Green = autotrophs, brown = protist heterotrophs, blue = dissolved nitrogen.

Insets show assumed community composition (dots) in size classes (increasing in size upwards), with lines indicating trophic relationships. Graphs show the fraction of total nitrogen (‘frac N’) in dissolved form (blue) and in each component of the size structured population, as total nitrogen (‘tot N’) increases. Heterophs are stacked on top of autotrophs and for each, where different size classes coexist, the smallest size class is at the bottom with progressively larger size classes stacked on top: (a) cyanobacteria only — nutrients are drawn down to limiting concentration and cyanobacterial population increases (until light limitation, not shown); (b) including phagotrophy from heterotrophic nanoflagellates limits cyanobacterial population size, allowing nutrient levels to rise; (c) this allows autotrophic nanoflagellates to coexist with cyanobacteria; (d) eukaryophagy limits population size of autotrophic nanoflagellates and allows larger size classes of eukaryotic phytoplankton (at high total nitrogen levels).

Figure 3.
Community size structure in an idealized steady-state microbial food chain model [52,111], for assumed evolutionary steps. Green = autotrophs, brown = protist heterotrophs, blue = dissolved nitrogen.

Insets show assumed community composition (dots) in size classes (increasing in size upwards), with lines indicating trophic relationships. Graphs show the fraction of total nitrogen (‘frac N’) in dissolved form (blue) and in each component of the size structured population, as total nitrogen (‘tot N’) increases. Heterophs are stacked on top of autotrophs and for each, where different size classes coexist, the smallest size class is at the bottom with progressively larger size classes stacked on top: (a) cyanobacteria only — nutrients are drawn down to limiting concentration and cyanobacterial population increases (until light limitation, not shown); (b) including phagotrophy from heterotrophic nanoflagellates limits cyanobacterial population size, allowing nutrient levels to rise; (c) this allows autotrophic nanoflagellates to coexist with cyanobacteria; (d) eukaryophagy limits population size of autotrophic nanoflagellates and allows larger size classes of eukaryotic phytoplankton (at high total nitrogen levels).

Close modal

A microbially dominated DOM recycling ecosystem would have had interesting implications for water column redox state [6]. Unlike the biological pump in the modern ocean (due to sinking POM), oxygen demand in the water column would be strongly influenced by advective and diffusive transport of DOM. This would still be able to drive anoxia in the open-ocean thermocline (see Figure 5f of [6]) and in shelf seas (Figure 4a), although less efficiently for a given nutrient level than particulate export. Shelf-sea environments would be expected to show strong latitudinal patterns in redox state, with permanent anoxia likely in permanently stratified regions at low-mid latitudes, and seasonal anoxia in seasonally stratifying high-latitudes shelves.

Scenarios for different shelf sea biogeochemical regimes through time, showing redox state after adjustment of P cycle (∼104 year) but before any adjustment of O2 cycle.

Figure 4.
Scenarios for different shelf sea biogeochemical regimes through time, showing redox state after adjustment of P cycle (∼104 year) but before any adjustment of O2 cycle.

Physical setting is a stratified shelf sea ∼100 m deep with surface waters separated from deeper waters by a sharp thermocline. DOM, dissolved organic matter (pale green). POM, particulate organic matter (dark green). Arrow width roughly represents magnitude of organic matter flux. We assume constant P input flux throughout hence P burial flux into sediments is identical throughout, but this is achieved at different [P] and through a different balance of processes over time. (a) Tonian ‘DOM world’ either dominated by cyanobacterial productivity (small pale green dots) or including small green algae (dark green circles) and small phagotrophic eukaryotes (brown ‘Pac-Men’), overlying benthic mats (pale green layer). Dashed arrows indicate uncertainty surrounding POM pathway(s) to sediments from sinking ‘marine snow’ and/or photoautotrophic mats. (b) Cryogenian world of eukaryotic algae (larger dark green dots) and eukaryophagy (larger brown ‘Pac-Men’) with a biological pump transferring POM to sediments. (c) Late Ediacaran world of sessile animals, including rangeomorph fronds (brown diamonds on stalks) and filter-feeding sponges (brown clouds) transferring POM to sediments. Their location on top of benthic mats creates a sharp redox boundary supporting phosphorite and authigenic carbonate deposition. (d) Cambrian world of mobile animals (brown splodges) bioturbating (and thus oxygenating) upper sediments, which lowers the Corg/P burial ratio enabling a smaller sedimentary POM flux to maintain the required P output flux.

Figure 4.
Scenarios for different shelf sea biogeochemical regimes through time, showing redox state after adjustment of P cycle (∼104 year) but before any adjustment of O2 cycle.

Physical setting is a stratified shelf sea ∼100 m deep with surface waters separated from deeper waters by a sharp thermocline. DOM, dissolved organic matter (pale green). POM, particulate organic matter (dark green). Arrow width roughly represents magnitude of organic matter flux. We assume constant P input flux throughout hence P burial flux into sediments is identical throughout, but this is achieved at different [P] and through a different balance of processes over time. (a) Tonian ‘DOM world’ either dominated by cyanobacterial productivity (small pale green dots) or including small green algae (dark green circles) and small phagotrophic eukaryotes (brown ‘Pac-Men’), overlying benthic mats (pale green layer). Dashed arrows indicate uncertainty surrounding POM pathway(s) to sediments from sinking ‘marine snow’ and/or photoautotrophic mats. (b) Cryogenian world of eukaryotic algae (larger dark green dots) and eukaryophagy (larger brown ‘Pac-Men’) with a biological pump transferring POM to sediments. (c) Late Ediacaran world of sessile animals, including rangeomorph fronds (brown diamonds on stalks) and filter-feeding sponges (brown clouds) transferring POM to sediments. Their location on top of benthic mats creates a sharp redox boundary supporting phosphorite and authigenic carbonate deposition. (d) Cambrian world of mobile animals (brown splodges) bioturbating (and thus oxygenating) upper sediments, which lowers the Corg/P burial ratio enabling a smaller sedimentary POM flux to maintain the required P output flux.

Close modal

With all oxygen demand used up at intermediate depths, deeper waters of the open ocean would be less prone to anoxia (Figure 5f of [6]), having their [O2] governed by the balance of supply from the atmosphere via high-latitude deep convection and (a smaller) demand from inorganic reductant input at mid-ocean ridges [6]. Any additional POM flux into deep waters could readily have driven them anoxic (Figure 5d of [6] and [41]). However, what little data we have from the truly deep Proterozoic ocean suggest that some oxygen was present [42,43], consistent with a DOM-dominated system without a significant biological pump [6].

Maintaining an oxygenated Proterozoic atmosphere (given tectonic reductant input) required greater than ∼25% of present-day Corg burial [17]. A key puzzle is; how did organic matter — including P and C — get out of the bottom of this DOM-cycling system in solid, sedimentary form? Much of it must have been derived from seafloor microbial mats. It seems unlikely that heterotrophic bacterial mats could extract significant DOM from the water column, relative to much larger water column bacterial populations. Instead, occasional aggregation of DOM into POM [44] and of bacterial cells into larger POM, which sunk as ‘marine snow’, seems a more plausible food source for heterotrophic mats. Additionally, autotrophic shallow water mats would have produced POM.

In this ‘DOM world’, balancing the P cycle would have demanded a larger ocean P inventory (Figure 4a) to drive the same P output, consistent with geochemical inferences of high P levels by the Cryogenian [29,45]. As ocean nutrient inventory determines the total amount of organic matter production in the ocean, a high P ocean would have been prone to anoxia at the depths where DOM was respired, particularly under lower early Neoproterozoic atmospheric pO2 [6].

Fossils indicate a diversification of heterotrophic eukaryotes in the late Tonian [4648], including testate amoebae and organic or siliceous scales, followed by the advent of foraminifera and ciliates in the Cryogenian [49] (Figure 1b). Recent (re)analysis of the biomarker record [34] indicates that steranes from eukaryotic algae, demosponges and rhizarians (a group which includes foraminifera and radiolarians) all first appear in the Cryogenian period (720–635 Ma), between the Sturtian and Marinoan ‘snowball Earth’ glaciations ∼660–640 Ma [50], and the sterane/hopane ratio increases, indicating the prevalence of eukaryotic over bacterial lipids [34] (Figure 1a).

The absence of a eukaryotic algal biomarker record prior to the Cryogenian has been attributed to nutrient scarcity [34,51]. However, in the modern ocean phagotrophy limits cyanobacterial population size allowing eukaryotic algae to coexist [52], where even in the most oligotrophic regions mixotrophic eukaryotes that consume bacteria compete effectively with prokaryotes [53]. Also, during Phanerozoic ocean anoxic events that deplete nitrogen, eukaryotes comprise a significant fraction of productivity [54]. Furthermore, any boost in phosphorus levels during [45] or after the Sturtian glaciation would be transient [1] hence cannot explain the irreversible biomarker transition.

Instead, we suggest two scenarios linking the biomarker record of eukaryotic algae [34] with an apparent escalation in protistan predation [55,56] (Figure 3). In the first scenario, eukaryotic bacterivory was, for some reason, ineffective prior to the Cryogenian (despite being plesiomorphic among eukaryotes), allowing prokaryotic autotrophs to competitively exclude larger eukaryotic algae, because smaller cells are more efficient at diffusion-limited nutrient uptake [52] (Figure 3a). Then, an increase in the effectiveness of eukaryotic bacteriophagy limited cyanobacterial population size (Figure 3b, as in the modern ocean), allowing surface nutrient levels to rise and creating a niche for autotrophic eukaryotes (Figure 3c), followed by a rapid escalation of eukaryophagy to exploit this new resource [57] (Figure 3d). Alternatively, the advent of eukaryophagy put a selection pressure for predation resistance on an earlier cryptic population of small heterotrophic, autotrophic and mixotrophic eukaryotes (Figure 3c), driving increases in size (Figure 3d), armour, and therefore sinking speed and preservation potential — leading to the formation of the algal biomarker record. By filtering out smaller cyanobacterial cells, the advent of sponges [5860] could also have provided a strong selection pressure for larger eukaryotic algal cells in shelf seas [1].

Thanks to their larger size and faster sinking rate, eukaryotes created a biological pump of POM [1]. The short-term effect would have been to increase oxygen demand in the deeper waters of shelf seas and the open ocean [6,61]. However, by creating an efficient particulate P removal flux to shelf sediments, it would have lowered ocean P content and corresponding O2 demand over ∼104 years, thus tending to oxygenate deeper waters of shelf seas (Figure 4b) and the open ocean [6]. This prediction is supported by redox proxy evidence of ocean oxygenation after the Sturtian [10,11,62]. The contribution of redox-sensitive P cycling feedback to this oxygenation is uncertain, but the partial oxygenation of ‘ferruginous’ background ocean waters (Figure 1c) apparently did not cause a major drop in the efficiency of P removal with Fe minerals [2729], as that would have rapidly counteracted further oxygenation. The oxygenation event is inferred to have reversed before the Marinoan [11], perhaps because partial shelf seafloor oxygenation produced a lower Corg to total phosphorus burial ratio [21], thus triggering a decline in atmospheric pO2.

The nature and timing (Figure 1a,b) of the first animals remains contested. Recent molecular phylogenetic studies support the common sense view that sponges (Porifera) are the sister group to all other animals [63,64]. Relaxed molecular clocks put the origin of crown-group demosponges 872–657 Ma (across studies) [65] consistent with biomarkers ∼660–640 Ma [5860], but put silicification and spicule production later at 648–616 Ma [66]. Hence, the gap to the first widely accepted fossil sponges in the early Cambrian ∼535 Ma [6769] could be partly due to poor preservation potential. One ∼600 Ma fossil [70] could also close this gap, if supported. A counter-view takes the fossil record at face value (rejecting the molecular clock and biomarker evidence) and argues sponges are not the basal animals and originated in the latest Ediacaran–Cambrian [68,69].

The first complex macroscopic body fossils of the Ediacara biota ∼570–540 Ma are interpreted as a mix of stem- and crown-group animals [71,72]. Early ‘rangeomorph’ fronds that stuck up from the deep, dark seafloor could have fed by osmotrophy [73] — enhanced in deep, low-flow regimes by the establishment of a ‘canopy flow’ regime [74] — facilitating uptake of dissolved nutrients. However, if we accept that the Ediacara biota were eukaryotic, then they would have had other feeding modes — phagocytosis — and their own source of motility — flagella — with which to create advection that can significantly enhance nutrient uptake [75]. Either way, fronds could have provided a spatially concentrated source of POM to sediments (e.g. on death), thus lowering DOM and P concentration in the water column [76].

Whenever actively water-pumping, filter-feeding sponges appeared, they would have altered the size structure of the water column ecosystem, ocean nutrient levels and redox state (Figure 4c). As well as efficiently filtering bacteria and POM, in modern (oligotrophic) coral reef settings, the ‘sponge loop’ [77] converts DOM to POM — forming an effective nutrient concentration and recycling system and a mechanism for transferring POM to sediments [78]. Sponge symbionts also sequester polyphosphate, which can trigger apatite formation, thus removing P to sediments [79]. Thus, when sponges arose they could have provided a significant direct pathway of C and P to sediments, lowering overall oxygen demand in the water column and thus tending to oxygenate it. Thectardis ∼565–555 Ma has been interpreted as a possible sponge [80], although others disagree [68,69]. Subsequently, there is geochemical evidence for the progressive expansion of siliceous sponges on the Yangtze Platform ∼550–525 Ma, and associated decrease in the DOC pool, enhanced Corg and phosphorus burial, and water column oxygenation [81].

The combination of an oxygenated shelf water column overlying a sediment surface still covered by mats (and sessile animals) produced a shelf-sea ecosystem structure unique to the Neoproterozoic, with a sharp sediment-surface redox gradient that would favour the formation of authigenic carbonate [82] and phosphorite [8385]. In modern environments, sulphide-oxidising bacterial mats (Thiomargarita, Beggiatoa) that bridge the water-sediment interface accumulate polyphosphate from oxic waters and utilise it under anoxic conditions, triggering apatite precipitation [86,87]. Neoproterozoic phosphorites are typically associated with stromatolites [88,89] and some contain filamentous microfossils that resemble modern sulphide-oxidising bacteria [90]. Phosphorites are associated with nearshore oxygen oases ∼610 Ma [91], then shift to greater depths and areal extent ∼570 Ma onwards (e.g. the Doushantuo Formation) [83,91].

The first mobile trace-makers that scratched across mat surfaces appeared ∼565 Ma [92], but did not significantly disrupt the ‘mat seal’ on the sediments, until the first burrowing animals evolved. Fine meiofaunal traces from 555 to 542 Ma have recently been described [93], and widespread burrows capable of significant sediment mixing slightly predate the Precambrian/Cambrian boundary [94].

Bioturbating animals are well-known ‘ecosystem engineers’ [95] that bring oxygen into contact with sediments, increase the turnover rate of Corg and O2 [96], suppressing sulfate reduction [97] and trapping phosphorus in iron oxides [98], while also increasing water exchange fluxes that release nitrogen and phosphorus to the water column [97]. It is hypothesised that by oxidising upper sediment layers, the evolution of bioturbators increased the sulfate inventory of the ocean [99], and by lowering (Corg/P)burial, initially removed phosphorus and oxygenated the ocean (∼104 years) (Figure 4d) and then lowered atmospheric pO2 and deoxygenated the ocean (∼107 years) restoring higher P levels [100]. What is unresolved is when these predicted effects became globally significant.

One view is that bioturbation increased in depth and intensity in the Cambrian ‘agronomic revolution’ in Stages 2–4 (∼530–510 Ma) [101,102], causing initial ocean oxygenation and then declining atmospheric pO2 and ocean deoxygenation over the next ∼20 Myr [100]. A counter view is that the development of bioturbation only became globally significant from the late Silurian ∼420 Ma onwards [103,104]. The argument depends on whether sediment P cycling responds linearly or non-linearly to increasing burrowing depth. Diagenetic modelling suggests the effects of burrowing animals are nonlinear and even shallow bioturbation significantly sequesters P [105].

These alternative hypotheses make distinct, testable predictions. Evidence that mid-depth waters of the Yangtze Platform oxygenated from Cambrian stages 2–4 has been used to question the early bioturbation model [106], but is actually consistent with the original predictions [100], which show that ocean oxygenation should accompany the initial onset of deep bioturbation (Stages 2–4), followed by a much slower deoxygenation (governed by the slow timescale of atmospheric pO2 decline). Wider evidence shows an ocean oxygenation event ∼520 Ma (broadly coincident with the ‘Cambrian explosion’) followed by deoxygenation [13,107,108].

The Cambrian evolution of large zooplankton would also have increased the efficiency of the biological pump [36], transferring organic matter to sediments, lowering the ocean P inventory and tending to oxygenate the ocean [1,107].

The Neoproterozoic–Cambrian transition was not unidirectional or driven solely by either rising atmospheric pO2 or evolutionary innovations. We describe a series of eukaryotic innovations that created and strengthened the biological pump of POM from the ocean to the sediments, with major consequences for the phosphorus, carbon and oxygen cycles. We suggest that each phase of eukaryote evolution tended to lower P levels and oxygenate the ocean on ∼104 year timescales, but by decreasing Corg/P burial ratios tended to lower atmospheric pO2 and deoxygenate the ocean again on ∼106 year timescales. Coupled with tectonic drivers, and biogeochemical and ecological feedback, potentially between alternate stable states [14], this could help explain the transient nature and ∼106 year duration of oceanic oxygenation events through the Cryogenian–Ediacaran–Cambrian.

Summary
  • The late Tonian ocean ∼750 Ma was dominated by rapid microbial cycling of DOM with elevated nutrient (P) levels due to inefficient organic matter removal to sediments.

  • We hypothesise that abrupt onset of the eukaryotic algal biomarker record in the Cryogenian ∼660–640 Ma was linked to an escalation of protozoan predation (eukaryophagy).

  • This ecological regime shift created a ‘biological pump’ of sinking POM, which transferred Corg and P to sediments.

  • The Late Ediacaran advent of sessile benthic animals on top of microbial mats increased the efficiency of Corg and P transfer to sediments, contributing to the deposition of phosphorites.

  • The Cambrian explosion of mobile burrowing animals broke the ‘mat seal’ on the upper sediments but by oxygenating them it enabled alternative means of efficient P retention.

  • Each phase of eukaryote evolution tended to lower P levels and oxygenate the ocean on ∼104 year timescales, but by decreasing Corg/P burial ratios, tended to lower atmospheric pO2 and deoxygenate the ocean again on ∼106 year timescales.

  • This can help explain the transient nature and ∼106 year duration of oceanic oxygenation events through the Cryogenian–Ediacaran–Cambrian periods.

Corg

organic carbon

DOM

dissolved organic matter

POM

particulate organic matter

This work was supported by the NERC-NSFC programme ‘Biosphere Evolution, Transitions and Resilience’ through grant NE/P013651/1.

The Authors declare that there are no competing interests associated with the manuscript.

1
Lenton
,
T.M.
,
Boyle
,
R.A.
,
Poulton
,
S.W.
,
Shields
,
G.A.
and
Butterfield
,
N.J.
(
2014
)
Co-evolution of eukaryotes and ocean oxygenation in the Neoproterozoic era
.
Nat. Geosci.
7
,
257
265
2
Shields-Zhou
,
G.
and
Och
,
L.
(
2011
)
The case for a Neoproterozoic Oxygenation Event: geochemical evidence and biological consequences
.
GSA Today
21
,
4
11
3
Och
,
L.M.
and
Shields-Zhou
,
G.A.
(
2012
)
The Neoproterozoic oxygenation event: environmental perturbations and biogeochemical cycling
.
Earth-Sci. Rev.
110
,
26
57
4
Lyons
,
T.W.
,
Reinhard
,
C.T.
and
Planavsky
,
N.J.
(
2014
)
The rise of oxygen in Earth's early ocean and atmosphere
.
Nature
506
,
307
315
5
Planavsky
,
N.J.
,
Reinhard
,
C.T.
,
Wang
,
X.
,
Thomson
,
D.
,
McGoldrick
,
P.
,
Rainbird
,
R.H.
et al
(
2014
)
Low Mid-Proterozoic atmospheric oxygen levels and the delayed rise of animals
.
Science
346
,
635
638
6
Lenton
,
T.M.
and
Daines
,
S.J.
(
2017
)
Biogeochemical transformations in the history of the ocean
.
Ann. Rev. Mar. Sci.
9
,
31
58
7
Lenton
,
T.M.
and
Watson
,
A.J.
(
2000
)
Redfield revisited: 1. Regulation of nitrate, phosphate, and oxygen in the ocean
.
Global Biogeochem. Cycles
14
,
225
248
8
Monteiro
,
F.M.
,
Pancost
,
R.D.
,
Ridgwell
,
A.
and
Donnadieu
,
Y.
(
2012
)
Nutrients as the dominant control on the spread of anoxia and euxinia across the Cenomanian-Turonian oceanic anoxic event (OAE2): model-data comparison
.
Paleoceanography
27
,
PA4209
9
Lenton
,
T.M.
,
Daines
,
S.J.
and
Mills
,
B.J.W.
(
2018
)
COPSE reloaded: an improved model of biogeochemical cycling over Phanerozoic time
.
Earth Sci. Rev.
178
,
1
28
10
Zhang
,
F.
,
Zhu
,
X.
,
Yan
,
B.
,
Kendall
,
B.
,
Peng
,
X.
,
Li
,
J.
et al
(
2015
)
Oxygenation of a Cryogenian ocean (Nanhua Basin, South China) revealed by pyrite Fe isotope compositions
.
Earth Planet. Sci. Lett.
429
,
11
19
11
Lau
,
K.V.
,
Macdonald
,
F.A.
,
Maher
,
K.
and
Payne
,
J.L.
(
2017
)
Uranium isotope evidence for temporary ocean oxygenation in the aftermath of the sturtian snowball earth
.
Earth Planet. Sci. Lett.
458
,
282
292
12
Hardisty
,
D.S.
,
Lu
,
Z.
,
Bekker
,
A.
,
Diamond
,
C.W.
,
Gill
,
B.C.
,
Jiang
,
G.
et al
(
2017
)
Perspectives on Proterozoic surface ocean redox from iodine contents in ancient and recent carbonate
.
Earth Planet. Sci. Lett.
463
,
159
170
13
Sahoo
,
S.K.
,
Planavsky
,
N.J.
,
Jiang
,
G.
,
Kendall
,
B.
,
Owens
,
J.D.
,
Wang
,
X.
et al
(
2016
)
Oceanic oxygenation events in the anoxic Ediacaran ocean
.
Geobiology
14
,
457
468
14
Handoh
,
I.C.
and
Lenton
,
T.M.
(
2003
)
Periodic mid-Cretaceous Oceanic Anoxic Events linked by oscillations of the phosphorus and oxygen biogeochemical cycles
.
Global Biogeochem. Cycles
17
,
1092
15
England
,
M.H.
(
1995
)
The age of water and ventilation timescales in a global ocean model
.
J. Phys. Oceanogr.
25
,
2756
2777
16
Wallmann
,
K.
(
2010
)
Phosphorus imbalance in the global ocean?
Global Biogeochem. Cycles
24
,
GB4030
17
Daines
,
S.J.
,
Mills
,
B.J.W.
and
Lenton
,
T.M.
(
2017
)
Atmospheric oxygen regulation at low Proterozoic levels by incomplete oxidative weathering of sedimentary organic carbon
.
Nat. Commun.
8
,
14379
18
Slomp
,
C.P.
and
Van Cappellen
,
P.
(
2007
)
The global marine phosphorus cycle: sensitivity to oceanic circulation
.
Biogeosciences
4
,
155
171
19
Burdige
,
D.J.
(
2007
)
Preservation of organic matter in marine sediments: controls, mechanisms, and an imbalance in sediment organic carbon budgets?
Chem. Rev.
107
,
467
485
20
Ingall
,
E.
and
Jahnke
,
R.
(
1994
)
Evidence for enhanced phosphorus regeneration from marine sediments overlain by oxygen depleted waters
.
Geochim. Cosmochim. Acta
58
,
2571
2575
21
Ingall
,
E.D.
,
Bustin
,
R.M.
and
Van Cappellen
,
P.
(
1993
)
Influence of water column anoxia on the burial and preservation of carbon and phosphorus in marine shales
.
Geochim. Cosmochim. Acta
57
,
303
316
22
Sommer
,
S.
,
Clemens
,
D.
,
Yücel
,
M.
,
Pfannkuche
,
O.
,
Hall
,
P.O.J.
,
Almroth-Rosell
,
E.
et al
(
2017
)
Major bottom water ventilation events do not significantly reduce basin-wide benthic N and P release in the eastern Gotland basin (Baltic Sea)
.
Front. Mar. Sci.
4
,
18
23
Van Cappellen
,
P.
and
Ingall
,
E.D.
(
1994
)
Benthic phosphorus regeneration, net primary production, and ocean anoxia: a model of the coupled marine biogeochemical cycles of carbon and phosphorus
.
Paleoceanography
9
,
677
692
24
Van Cappellen
,
P.
and
Ingall
,
E.D.
(
1996
)
Redox stabilization of the atmosphere and oceans by phosphorus-limited marine productivity
.
Science
271
,
493
496
25
Poulton
,
S.W.
and
Canfield
,
D.E.
(
2011
)
Ferruginous conditions: a dominant feature of the ocean through Earth's history
.
Elements
7
,
107
112
26
Zegeye
,
A.
,
Bonneville
,
S.
,
Benning
,
L.G.
,
Sturm
,
A.
,
Fowle
,
D.A.
,
Jones
,
C.A.
et al
(
2012
)
Green rust formation controls nutrient availability in a ferruginous water column
.
Geology
40
,
599
602
27
Derry
,
L.A.
(
2015
)
Causes and consequences of mid-Proterozoic anoxia
.
Geophys. Res. Lett.
42
,
8538
8546
28
Laakso
,
T.A.
and
Schrag
,
D.P.
(
2014
)
Regulation of atmospheric oxygen during the Proterozoic
.
Earth Planet. Sci. Lett.
388
,
81
91
29
Reinhard
,
C.T.
,
Planavsky
,
N.J.
,
Gill
,
B.C.
,
Ozaki
,
K.
,
Robbins
,
L.J.
,
Lyons
,
T.W.
et al
(
2017
)
Evolution of the global phosphorus cycle
.
Nature
541
,
386
389
30
Parfrey
,
L.W.
,
Lahr
,
D.J.G.
,
Knoll
,
A.H.
and
Katz
,
L.A.
(
2011
)
Estimating the timing of early eukaryotic diversification with multigene molecular clocks
.
Proc. Natl Acad. Sci. U.S.A.
108
,
13624
13629
31
Butterfield
,
N.J.
(
2000
)
Bangiomorpha pubescens n. gen., n. sp.: implications for the evolution of sex, multicellularity, and the Mesoproterozoic/Neoproterozoic radiation of eukaryotes
.
Paleobiology
26
,
386
404
32
Sánchez-Baracaldo
,
P.
,
Raven
,
J.A.
,
Pisani
,
D.
and
Knoll
,
A.H.
(
2017
)
Early photosynthetic eukaryotes inhabited low-salinity habitats
.
Proc. Natl Acad. Sci. U.S.A.
114
,
E7737
E7745
33
Butterfield
,
N.J.
(
2015
)
Proterozoic photosynthesis - a critical review
.
Palaeontology
58
,
953
972
34
Brocks
,
J.J.
,
Jarrett
,
A.J.M.
,
Sirantoine
,
E.
,
Hallmann
,
C.
,
Hoshino
,
Y.
and
Liyanage
,
T.
(
2017
)
The rise of algae in Cryogenian oceans and the emergence of animals
.
Nature
548
,
578
581
35
Brocks
,
J.J.
,
Jarrett
,
A.J.M.
,
Sirantoine
,
E.
,
Kenig
,
F.
,
Moczydłowska
,
M.
,
Porter
,
S.
et al
(
2016
)
Early sponges and toxic protists: possible sources of cryostane, an age diagnostic biomarker antedating Sturtian Snowball Earth
.
Geobiology
14
,
129
149
36
Logan
,
G.A.
,
Hayes
,
J.M.
,
Hieshima
,
G.B.
and
Summons
,
R.E.
(
1995
)
Terminal Proterozoic reorganization of biogeochemical cycles
.
Nature
376
,
53
56
37
Pawlowska
,
M.M.
,
Butterfield
,
N.J.
and
Brocks
,
J.J.
(
2013
)
Lipid taphonomy in the Proterozoic and the effect of microbial mats on biomarker preservation
.
Geology
41
,
103
106
38
Hansell
,
D.A.
,
Carlson
,
C.A.
,
Repeta
,
D.J.
and
Schlitzer
,
R.
(
2009
)
Dissolved organic matter in the ocean: a controversy stimulates new insights
.
Oceanography
22
,
202
211
39
Jiao
,
N.
,
Herndl
,
G.J.
,
Hansell
,
D.A.
,
Benner
,
R.
,
Kattner
,
G.
,
Wilhelm
,
S.W.
et al
(
2010
)
Microbial production of recalcitrant dissolved organic matter: long-term carbon storage in the global ocean
.
Nat. Rev. Microbiol.
8
,
593
599
40
Rothman
,
D.H.
,
Hayes
,
J.M.
and
Summons
,
R.E.
(
2003
)
Dynamics of the Neoproterozoic carbon cycle
.
Proc. Natl Acad. Sci. U.S.A.
100
,
8124
8129
41
Reinhard
,
C.T.
,
Planavsky
,
N.J.
,
Olson
,
S.L.
,
Lyons
,
T.W.
and
Erwin
,
D.H.
(
2016
)
Earth's oxygen cycle and the evolution of animal life
.
Proc. Natl Acad. Sci. U.S.A.
113
,
8933
8938
42
Slack
,
J.F.
,
Grenne
,
T.
and
Bekker
,
A.
(
2009
)
Seafloor-hydrothermal Si-Fe-Mn exhalites in the Pecos greenstone belt, New Mexico, and the redox state of ca. 1720 Ma deep seawater
.
Geosphere
5
,
302
314
43
Slack
,
J.F.
,
Grenne
,
T.
,
Bekker
,
A.
,
Rouxel
,
O.J.
and
Lindberg
,
P.A.
(
2007
)
Suboxic deep seawater in the late Paleoproterozoic: evidence from hematitic chert and iron formation related to seafloor-hydrothermal sulfide deposits, central Arizona, USA
.
Earth Planet. Sci. Lett.
255
,
243
256
44
Engel
,
A.
,
Thoms
,
S.
,
Riebesell
,
U.
,
Rochelle-Newall
,
E.
and
Zondervan
,
I.
(
2004
)
Polysaccharide aggregation as a potential sink of marine dissolved organic carbon
.
Nature
428
,
929
932
45
Planavsky
,
N.J.
,
Rouxel
,
O.J.
,
Bekker
,
A.
,
Lalonde
,
S.V.
,
Konhauser
,
K.O.
,
Reinhard
,
C.T.
et al
(
2010
)
The evolution of the marine phosphate reservoir
.
Nature
467
,
1088
1090
46
Porter
,
S.M.
and
Knoll
,
A.H.
(
2000
)
Testate amoebae in the Neoproterozoic Era: evidence from vase-shaped microfossils in the Chuar Group, Grand Canyon
.
Paleobiology
26
,
360
385
47
Cohen
,
P.A.
,
Schopf
,
J.W.
,
Butterfield
,
N.J.
,
Kudryavtsev
,
A.B.
and
Macdonald
,
F.A.
(
2011
)
Phosphate biomineralization in mid-Neoproterozoic protists
.
Geology
39
,
539
542
48
Javaux
,
E.
(
2011
) Early eukaryotes in Precambrian oceans. In
Origins and Evolution of Life: An Astrobiological Perspective
(
Gargaud
,
M.
,
Lopez-Garcia
,
P.
and
Martin
,
H.
, eds), pp.
414
449
,
Cambridge University Press
,
Cambridge
49
Cohen
,
P.A.
and
Macdonald
,
F.A.
(
2015
)
The proterozoic record of eukaryotes
.
Paleobiology
41
,
610
632
50
Rooney
,
A.D.
,
Macdonald
,
F.A.
,
Strauss
,
J.V.
,
Dudás
,
F.Ö.
,
Hallmann
,
C.
and
Selby
,
D.
(
2014
)
Re-Os geochronology and coupled Os-Sr isotope constraints on the Sturtian snowball Earth
.
Proc. Natl Acad. Sci. U.S.A.
111
,
51
56
51
Knoll
,
A.H.
(
2017
)
Food for early animal evolution
.
Nature
548
,
528
530
52
Armstrong
,
R.A.
(
1994
)
Grazing limitation and nutrient limitation in marine ecosystems: steady state solutions of an ecosystem model with multiple food chains
.
Limnol. Oceanogr.
39
,
597
608
53
Hartmann
,
M.
,
Grob
,
C.
,
Tarran
,
G.A.
,
Martin
,
A.P.
,
Burkill
,
P.H.
,
Scanlan
,
D.J.
et al
(
2012
)
Mixotrophic basis of Atlantic oligotrophic ecosystems
.
Proc. Natl Acad. Sci. U.S.A.
109
,
5756
5760
54
Higgins
,
M.B.
,
Robinson
,
R.S.
,
Husson
,
J.M.
,
Carter
,
S.J.
and
Pearson
,
A.
(
2012
)
Dominant eukaryotic export production during ocean anoxic events reflects the importance of recycled NH4+
.
Proc. Natl Acad. Sci. U.S.A.
109
,
2269
2274
55
Knoll
,
A.H.
(
2014
)
Paleobiological perspectives on early eukaryotic evolution
.
Cold Spring Harbor Perspect. Biol.
6
,
a016121
56
Porter
,
S.
(
2011
)
The rise of predators
.
Geology
39
,
607
608
57
Knoll
,
A.H.
and
Follows
,
M.J.
(
2016
)
A bottom-up perspective on ecosystem change in Mesozoic oceans
.
Proc. Royal Soc. B: Biol. Sci.
283
,
20161755
58
Love
,
G.D.
,
Grosjean
,
E.
,
Stalvies
,
C.
,
Fike
,
D.A.
,
Grotzinger
,
J.P.
,
Bradley
,
A.S.
et al
(
2009
)
Fossil steroids record the appearance of Demospongiae during the Cryogenian period
.
Nature
457
,
718
721
59
Gold
,
D.A.
,
Grabenstatter
,
J.
,
de Mendoza
,
A.
,
Riesgo
,
A.
,
Ruiz-Trillo
,
I.
and
Summons
,
R.E.
(
2016
)
Sterol and genomic analyses validate the sponge biomarker hypothesis
.
Proc. Natl Acad. Sci. U.S.A.
113
,
2684
2689
60
Love
,
G.D.
and
Summons
,
R.E.
(
2015
)
The molecular record of Cryogenian sponges - a response to Antcliffe (2013)
.
Palaeontology
58
,
1131
1136
61
Meyer
,
K.M.
,
Ridgwell
,
A.
and
Payne
,
J.L.
(
2016
)
The influence of the biological pump on ocean chemistry: implications for long-term trends in marine redox chemistry, the global carbon cycle, and marine animal ecosystems
.
Geobiology
14
,
207
219
62
Kunzmann
,
M.
,
Gibson
,
T.M.
,
Halverson
,
G.P.
,
Hodgskiss
,
M.S.W.
,
Bui
,
T.H.
,
Carozza
,
D.A.
et al
(
2017
)
Iron isotope biogeochemistry of Neoproterozoic marine shales
.
Geochim. Cosmochim. Acta
209
,
85
105
63
Simion
,
P.
,
Philippe
,
H.
,
Baurain
,
D.
,
Jager
,
M.
,
Richter
,
D.J.
,
Di Franco
,
A.
et al
(
2017
)
A large and consistent phylogenomic dataset supports sponges as the sister group to all other animals
.
Curr. Biol.
27
,
958
967
64
Feuda
,
R.
,
Dohrmann
,
M.
,
Pett
,
W.
,
Philippe
,
H.
,
Rota-Stabelli
,
O.
,
Lartillot
,
N.
et al
(
2017
)
Improved modeling of compositional heterogeneity supports sponges as sister to all other animals
.
Curr. Biol.
27
,
3864
3870.e4
65
Schuster
,
A.
,
Vargas
,
S.
,
Knapp
,
I.
,
Pomponi
,
S.A.
,
Toonen
,
R.J.
,
Erpenbeck
,
D.
et al
(
2017
)
Divergence times in demosponges (Porifera): first insights from new mitogenomes and the inclusion of fossils in a birth-death clock model
.
bioRxiv
66
Ma
,
J.-Y.
and
Yang
,
Q.
(
2016
)
Early divergence dates of demosponges based on mitogenomics and evaluated fossil calibrations
.
Palaeoworld
25
,
292
302
67
Botting
,
J.P.
,
Cárdenas
,
P.
and
Peel
,
J.S.
(
2015
)
A crown-group demosponge from the early Cambrian Sirius Passet Biota, North Greenland
.
Palaeontology
58
,
35
43
68
Antcliffe
,
J.B.
,
Callow
,
R.H.T.
and
Brasier
,
M.D.
(
2014
)
Giving the early fossil record of sponges a squeeze
.
Biol. Rev.
89
,
972
1004
69
Botting
,
J.P.
and
Muir
,
L.A.
(
2018
)
Early sponge evolution: a review and phylogenetic framework
.
Palaeoworld
27
,
1
29
70
Yin
,
Z.
,
Zhu
,
M.
,
Davidson
,
E.H.
,
Bottjer
,
D.J.
,
Zhao
,
F.
and
Tafforeau
,
P.
(
2015
)
Sponge grade body fossil with cellular resolution dating 60 Myr before the Cambrian
.
Proc. Natl Acad. Sci. U.S.A.
112
,
E1453
E1460
71
Droser
,
M.L.
,
Tarhan
,
L.G.
and
Gehling
,
J.G.
(
2017
)
The rise of animals in a changing environment: global ecological innovation in the late Ediacaran
.
Annu. Rev. Earth Planet. Sci.
45
,
593
617
72
Muscente
,
A.D.
,
Boag
,
T.H.
,
Bykova
,
N.
and
Schiffbauer
,
J.D.
(
2018
)
Environmental disturbance, resource availability, and biologic turnover at the dawn of animal life
.
Earth-Science Reviews.
177
,
248
264
73
Laflamme
,
M.
,
Xiao
,
S.
and
Kowalewski
,
M.
(
2009
)
Osmotrophy in modular Ediacara organisms
.
Proc. Natl Acad. Sci. U.S.A.
106
,
14438
14443
74
Ghisalberti
,
M.
,
Gold
,
D.A.
,
Laflamme
,
M.
,
Clapham
,
M.E.
,
Narbonne
,
G.M.
,
Summons
,
R.E.
et al
(
2014
)
Canopy flow analysis reveals the advantage of size in the oldest communities of multicellular eukaryotes
.
Curr. Biol.
24
,
305
309
75
Solari
,
C.A.
,
Ganguly
,
S.
,
Kessler
,
J.O.
,
Michod
,
R.E.
and
Goldstein
,
R.E.
(
2006
)
Multicellularity and the functional interdependence of motility and molecular transport
.
Proc. Natl Acad. Sci. U.S.A.
103
,
1353
1358
76
Budd
,
G.E.
and
Jensen
,
S.
(
2017
)
The origin of the animals and a ‘Savannah’ hypothesis for early bilaterian evolution
.
Biol. Rev.
92
,
446
473
77
de Goeij
,
J.M.
,
van Oevelen
,
D.
,
Vermeij
,
M.J.A.
,
Osinga
,
R.
,
Middelburg
,
J.J.
,
de Goeij
,
A.F.P.M.
et al
(
2013
)
Surviving in a marine desert: the sponge loop retains resources within coral reefs
.
Science
342
,
108
110
78
Middelburg
,
J.J.
(
2018
)
Reviews and syntheses: to the bottom of carbon processing at the seafloor
.
Biogeosciences
15
,
413
427
79
Zhang
,
F.
,
Blasiak
,
L.C.
,
Karolin
,
J.O.
,
Powell
,
R.J.
,
Geddes
,
C.D.
and
Hill
,
R.T.
(
2015
)
Phosphorus sequestration in the form of polyphosphate by microbial symbionts in marine sponges
.
Proc. Natl Acad. Sci. U.S.A.
112
,
4381
4386
80
Sperling
,
E.A.
,
Peterson
,
K.J.
and
Laflamme
,
M.
(
2011
)
Rangeomorphs, Thectardis (Porifera?) and dissolved organic carbon in the Ediacaran oceans
.
Geobiology
9
,
24
33
81
Tatzel
,
M.
,
von Blanckenburg
,
F.
,
Oelze
,
M.
,
Bouchez
,
J.
and
Hippler
,
D.
(
2017
)
Late Neoproterozoic seawater oxygenation by siliceous sponges
.
Nat. Commun.
8
,
621
82
Cui
,
H.
,
Kaufman
,
A.J.
,
Xiao
,
S.
,
Zhou
,
C.
and
Liu
,
X.-M.
(
2017
)
Was the Ediacaran Shuram Excursion a globally synchronized early diagenetic event? Insights from methane-derived authigenic carbonates in the uppermost Doushantuo Formation, South China
.
Chem. Geol.
450
,
59
80
83
Cui
,
H.
,
Xiao
,
S.
,
Zhou
,
C.
,
Peng
,
Y.
,
Kaufman
,
A.J.
and
Plummer
,
R.E.
(
2016
)
Phosphogenesis associated with the Shuram Excursion: petrographic and geochemical observations from the Ediacaran Doushantuo Formation of South China
.
Sediment. Geol.
341
,
134
146
84
Crosby
,
C.H.
and
Bailey
,
J.V.
(
2012
)
The role of microbes in the formation of modern and ancient phosphatic mineral deposits
.
Front. Microbiol.
3
,
241
85
Omelon
,
S.
,
Ariganello
,
M.
,
Bonucci
,
E.
,
Grynpas
,
M.
and
Nanci
,
A.
(
2013
)
A review of phosphate mineral nucleation in biology and geobiology
.
Calcif. Tissue Int.
93
,
382
396
86
Goldhammer
,
T.
,
Brüchert
,
V.
,
Ferdelman
,
T.G.
and
Zabel
,
M.
(
2010
)
Microbial sequestration of phosphorus in anoxic upwelling sediments
.
Nat. Geosci.
3
,
557
561
87
Schulz
,
H.N.
and
Schulz
,
H.D.
(
2005
)
Large sulfur bacteria and the formation of phosphorite
.
Science
307
,
416
418
88
Papineau
,
D.
(
2010
)
Global biogeochemical changes at both ends of the proterozoic: insights from phosphorites
.
Astrobiology
10
,
165
181
89
Gómez-Peral
,
L.E.
,
Kaufman
,
A.J.
and
Poiré
,
D.G.
(
2014
)
Paleoenvironmental implications of two phosphogenic events in Neoproterozoic sedimentary successions of the Tandilia System, Argentina
.
Precambrian Res.
252
,
88
106
90
Bailey
,
J.V.
,
Corsetti
,
F.A.
,
Greene
,
S.E.
,
Crosby
,
C.H.
,
Liu
,
P.
and
Orphan
,
V.J.
(
2013
)
Filamentous sulfur bacteria preserved in modern and ancient phosphatic sediments: implications for the role of oxygen and bacteria in phosphogenesis
.
Geobiology
11
,
397
405
91
Drummond
,
J.B.R.
,
Pufahl
,
P.K.
,
Porto
,
C.G.
and
Carvalho
,
M.
(
2015
)
Neoproterozoic peritidal phosphorite from the Sete Lagoas Formation (Brazil) and the Precambrian phosphorus cycle
.
Sedimentology
62
,
1978
2008
92
Liu
,
A.G.
,
Mcllroy
,
D.
and
Brasier
,
M.D.
(
2010
)
First evidence for locomotion in the Ediacara biota from the 565 Ma Mistaken Point Formation, Newfoundland
.
Geology
38
,
123
126
93
Parry
,
L.A.
,
Boggiani
,
P.C.
,
Condon
,
D.J.
,
Garwood
,
R.J.
,
de Leme
,
J..M.
,
McIlroy
,
D.
et al
(
2017
)
Ichnological evidence for meiofaunal bilaterians from the terminal Ediacaran and earliest Cambrian of Brazil
.
Nat. Ecol. Evol.
1
,
1455
1464
94
Jensen
,
S.
,
Saylor
,
B.Z.
,
Gehling
,
J.G.
and
Germs
,
G.J.B.
(
2000
)
Complex trace fossils from the terminal Proterozoic of Namibia
.
Geology
28
,
143
146
95
Mermillod-Blondin
,
F.
and
Rosenberg
,
R.
(
2006
)
Ecosystem engineering: the impact of bioturbation on biogeochemical processes in marine and freshwater benthic habitats
.
Aquat. Sci.
68
,
434
442
96
Kristensen
,
E.
(
2001
)
Impact of polychaetes (Nereis spp. and Arenicola marina) on carbon biogeochemistry in coastal marine sediments
. Geochem. Trans.
2
,
92
97
Mermillod-Blondin
,
F.
,
Rosenberg
,
R.
,
François-Carcaillet
,
F.
,
Norling
,
K.
and
Mauclaire
,
L.
(
2004
)
Influence of bioturbation by three benthic infaunal species on microbial communities and biogeochemical processes in marine sediment
.
Aquat. Microb. Ecol.
36
,
271
284
98
Zhang
,
L.
,
Gu
,
X.
,
Fan
,
C.
,
Shang
,
J.
,
Shen
,
Q.
,
Wang
,
Z.
et al
(
2010
)
Impact of different benthic animals on phosphorus dynamics across the sediment-water interface
.
J. Environ. Sci.
22
,
1674
1682
99
Canfield
,
D.E.
and
Farquhar
,
J.
(
2009
)
Animal evolution, bioturbation, and the sulfate concentration of the oceans
.
Proc. Natl Acad. Sci. U.S.A.
106
,
8123
8127
100
Boyle
,
R.A.
,
Dahl
,
T.W.
,
Dale
,
A.W.
,
Shields-Zhou
,
G.A.
,
Zhu
,
M.
,
Brasier
,
M.D.
et al
(
2014
)
Stabilization of the coupled oxygen and phosphorus cycles by the evolution of bioturbation
.
Nat. Geosci.
7
,
671
676
101
Mángano
,
M.G.
and
Buatois
,
L.A.
(
2017
)
The Cambrian revolutions: trace-fossil record, timing, links and geobiological impact
.
Earth-Sci. Rev.
173
,
96
108
102
Zhang
,
L.-J.
,
Qi
,
Y.-A.
,
Buatois
,
L.A.
,
Mángano
,
M.G.
,
Meng
,
Y.
and
Li
,
D.
(
2017
)
The impact of deep-tier burrow systems in sediment mixing and ecosystem engineering in early Cambrian carbonate settings
.
Sci. Rep.
7
,
45773
103
Tarhan
,
L.G.
and
Droser
,
M.L.
(
2014
)
Widespread delayed mixing in early to middle Cambrian marine shelfal settings
.
Palaeogeogr. Palaeoclimatol. Palaeoecol.
399
,
310
322
104
Tarhan
,
L.G.
,
Droser
,
M.L.
,
Planavsky
,
N.J.
and
Johnston
,
D.T.
(
2015
)
Protracted development of bioturbation through the early Palaeozoic Era
.
Nat. Geosci.
8
,
865
869
105
Dale
,
A.W.
,
Boyle
,
R.A.
,
Lenton
,
T.M.
,
Ingall
,
E.D.
and
Wallmann
,
K.
(
2016
)
A model for microbial phosphorus cycling in bioturbated marine sediments: significance for phosphorus burial in the early Paleozoic
.
Geochim. Cosmochim. Acta
189
,
251
268
106
Li
,
C.
,
Jin
,
C.
,
Planavsky
,
N.J.
,
Algeo
,
T.J.
,
Cheng
,
M.
,
Yang
,
X.
et al
(
2017
)
Coupled oceanic oxygenation and metazoan diversification during the early–middle Cambrian?
Geology
45
,
743
746
107
Dahl
,
T.W.
,
Connelly
,
J.N.
,
Kouchinsky
,
A.
,
Gill
,
B.C.
,
Månsson
,
S.F.
and
Bizzarro
,
M.
(
2017
)
Reorganisation of Earth's biogeochemical cycles briefly oxygenated the oceans 520 Myr ago
.
Geochem. Perspect. Lett.
3
,
210
220
108
Chen
,
X.
,
Ling
,
H.-F.
,
Vance
,
D.
,
Shields-Zhou
,
G.A.
,
Zhu
,
M.
,
Poulton
,
S.W.
et al
(
2015
)
Rise to modern levels of ocean oxygenation coincided with the Cambrian radiation of animals
.
Nat. Commun.
6
,
7142
109
Canfield
,
D.E.
,
Poulton
,
S.W.
,
Knoll
,
A.H.
,
Narbonne
,
G.M.
,
Ross
,
G.
,
Goldberg
,
T.
et al
(
2008
)
Ferruginous conditions dominated later neoproterozoic deep-water chemistry
.
Science
321
,
949
952
110
Canfield
,
D.E.
,
Poulton
,
S.W.
and
Narbonne
,
G.M.
(
2007
)
Late-Neoproterozoic deep-ocean oxygenation and the rise of animal life
.
Science
315
,
92
95
111
Ward
,
B.A.
,
Dutkiewicz
,
S.
and
Follows
,
M.J.
(
2014
)
Modelling spatial and temporal patterns in size-structured marine plankton communities: top–down and bottom–up controls
.
J. Plankton Res.
36
,
31
47
This is an open access article published by Portland Press Limited on behalf of the Biochemical Society and distributed under the Creative Commons Attribution License 4.0 (CC BY).