Hostname: page-component-848d4c4894-x24gv Total loading time: 0 Render date: 2024-05-16T19:02:07.048Z Has data issue: false hasContentIssue false

Constructional and functional anatomy of Ediacaran rangeomorphs

Published online by Cambridge University Press:  03 August 2020

Nicholas J Butterfield*
Affiliation:
Department of Earth Sciences, University of Cambridge, Cambridge, UKCB2 3EQ
*
Author for correspondence: Nicholas J Butterfield, Email: njb1005@cam.ac.uk
Rights & Permissions [Opens in a new window]

Abstract

Ediacaran rangeomorphs were the first substantially macroscopic organisms to appear in the fossil record, but their underlying biology remains problematic. Although demonstrably heterotrophic, their current interpretation as osmotrophic consumers of dissolved organic carbon (DOC) is incompatible with the inertial (high Re) and advective (high Pe) fluid dynamics accompanying macroscopic length scales. The key to resolving rangeomorph feeding and physiology lies in their underlying construction. Taphonomic analysis of three-dimensionally preserved Charnia from the White Sea identifies the presence of large, originally water-filled compartments that served both as a hydrostatic exoskeleton and semi-isolated digestion chambers capable of processing recalcitrant substrates, most likely in conjunction with a resident microbiome. At the same time, the hydrodynamically exposed outer surface of macroscopic rangeomorphs would have dramatically enhanced both gas exchange and food delivery. A bag-like epithelium filled with transiently circulated seawater offers an exceptionally efficient means of constructing a simple, DOC-consuming, multicellular heterotroph. Such a body plan is broadly comparable to that of anthozoan cnidarians, minus such derived features as muscle, tentacles and a centralized mouth. Along with other early bag-like fossils, rangeomorphs can be reliably identified as total-group eumetazoans, potentially colonial stem-group cnidarians.

Type
Original 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 in any medium, provided the original work is properly cited.
Copyright
© The Author(s), 2020. Published by Cambridge University Press

1. Introduction

Rangeomorphs were conspicuous members of the Ediacaran biosphere, present from roughly the end of the Gaskiers glaciation through to the beginning of the Cambrian period (c. 575–540 Ma). As the first substantial instances of large complex organisms in the fossil record, they mark a key transition in geobiological history, presaging the first appearance of unambiguous animals at c. 555 Ma. Even so, there is little consensus on the phylogenetic affiliations of rangeomorphs, with interpretations ranging from total-group cnidarians, ctenophores or sponges, to stem-group (eu)metazoans, or possibly an entirely unrelated lineage of multicellular macroscopic eukaryotes (Xiao & Laflamme, Reference Xiao and Laflamme2009). The problem with these particular fossils is not just their taxonomic placement, however, but a fundamentally deeper lack of understanding of how they were constructed, and how they worked as organisms.

Physiologically, macroscopic organisms work in much the same way as their microscopic counterparts, but with the added allometric challenges of conveying resources to internalized tissues and supporting the associated mass. Because of the fundamental reductions of surface area to volume (SA:V) that accompany increased body size, the evolution of large three-dimensional (3D) organisms necessarily involved major anatomical innovations. Even at modern levels of oxygen, for example, the maximum diameter of aerobic organisms lacking some sort of differentiated circulatory or respiratory apparatus is less than 2 mm (Catling et al. Reference Catling, Glein, Zahnle and McKay2005). There are, however, significant advantages to large body size, not least environmental buffering, systematic decreases in mass-specific metabolic demand (Glazier, Reference Glazier2006; DeLong et al. Reference DeLong, Okie, Moses, Sibly and Brown2010), and the emergence of scale-dependent mechanical, chemical and hydrodynamic properties (Sebens, Reference Sebens1987; Koehl, Reference Koehl1996; Hurd, Reference Hurd2000; Solari et al. Reference Solari, Kessler and Goldstein2007; Guizien & Ghisalberti, Reference Guizien, Ghisalberti, Rossi, Bramanti, Gori and Orejas Saco del Valle2016). Simply as a consequence of length scale (L) and background fluid velocity (U), for example, large organisms operate in a world of elevated Reynolds numbers where movement is dominated by inertial rather than viscous forces (Re = UL/v), and at elevated Péclet numbers where material exchange is dominated by advection rather than molecular diffusion (Pe = UL/D) (where v is kinematic viscosity and D is rate of diffusion). Such dynamics offer fundamentally enhanced levels of motility, feeding and gas exchange to macroscopic organisms (Butterfield, Reference Butterfield2018), provided the underlying issues of construction and resource distribution can be addressed.

Rangeomorphs are characterized by a broadly frond-like habit, built around a centimetre-scale branching element that exhibits self-similarity over three or four ‘fractal’ levels (Narbonne, Reference Narbonne2004; Brasier et al. Reference Brasier, Antcliffe and Liu2012; Hoyal Cuthill & Conway Morris, Reference Hoyal Cuthill and Conway Morris2014). Various arrangements of these elements, often in concert with basal holdfasts, elevating stalks or interconnecting rods, gave rise to a significant range of larger-scale forms including bi-terminal recliners (e.g. Fractofusus), unstalked unifoliate fronds (e.g. Charnia), stalked unifoliate fronds (e.g. Avalofractus), unstalked multifoliate fronds (e.g. Bradgatia), stalked multifoliate fronds (e.g. Rangea) and compound fronds (e.g. Hapsidophyllas) (Fig. 1). Although originally interpreted as macroscopic algae (Ford, Reference Ford1958), the abundant in situ preservation of rangeomorphs in deeper-water strata of Avalonia has convincingly ruled out a photosynthetic habit (Wood et al. Reference Wood, Dalrymple, Narbonne, Gehling and Clapham2003); and although some forms look superficially like modern sea-pens, their distinct styles of construction and growth rule out any direct affiliation to extant octocoral cnidarians (Seilacher, Reference Seilacher1989; Antcliffe & Brasier, Reference Antcliffe and Brasier2007).

Fig. 1. Rangeomorph taxa illustrating the characteristic fractal-like branching and diversity of overall form. (a) Charnia masoni, type specimen, from Charnwood Forest, UK. (b) Rangea schneiderhoehni, type specimen, from Namibia. (c) Hapsidophyllas flexibilis, from SE Newfoundland. (d) Fractofusus misrae from SE Newfoundland. (e) Bradgatia sp. from SE Newfoundland. Scale bar: (a, e) 2 cm; (b) 1.5 cm; (c) 4 cm; and (d) 3 cm. Photo credits: (a) Phil Wilby; (b) Dima Grazhdankin; (c) Olga Zhaxybayeva; (d) Alex Liu; and (e) Jean-Bernard Caron.

Because of their anchored and generally elevated habit, rangeomorphs have traditionally been interpreted as microphagous suspension feeders, ecologically analogous to sponges or anthozoan cnidarians (Jenkins, Reference Jenkins1985). Unlike these living forms, however, rangeomorphs appear to lack tentacles, openings or any other feeding-related features, even in specimens preserving detail on a scale of tens of micrometres (Narbonne, Reference Narbonne2004). As such, it has been widely assumed that nutrient uptake took place on the outside of the organism, after the manner of osmotrophic bacteria or fungi (McMenamin, Reference McMenamin1993; Laflamme et al. Reference Laflamme, Xiao and Kowalewski2009; Sperling et al. Reference Sperling, Peterson and Laflamme2011). Certainly the characteristic ‘fractal’ branching would have increased the proportion of exposed surface area on which this might have occurred, but it remains to be demonstrated that rangeomorphs could actually feed in this fashion (Liu et al. Reference Liu, Kenchington and Mitchell2015).

1.a. Osmotrophy and dissolved organic carbon: a primer

The most immediate issue arising from the osmotrophy model for rangeomorphs is the imprecise, often inconsistent, use of the term itself. In its most general sense, osmotrophy is simply the process by which dissolved substrates are passed across cell membranes (Jumars et al. Reference Jumars, Deming, Hill, Karp-Boss, Yager and Dade1993; Karp-Boss et al. Reference Karp-Boss, Boss and Jumars1996; Thingstad et al. Reference Thingstad, Øvreås, Egge, Løvdal and Heldal2005). At least implicitly, it is limited to external, environmentally exposed surfaces, which usefully distinguishes it from otherwise similar processes of internalized uptake (e.g. in eumetazoan guts and the food vacuoles of phagocytizing protozoans). The textbook exemplars of osmotrophic feeding – heterotrophic prokaryotes and fungi – are also ‘external digesters’, where organisms actively excrete hydrolytic enzymes and recover the digested products. There are risks to this type of feeding, however, most obviously through the dispersive loss of exo-enzymes and product in aqueous environments, but also their exploitation by unrelated or non-contributing organisms (Jumars et al. Reference Jumars, Deming, Hill, Karp-Boss, Yager and Dade1993; Karp-Boss et al. Reference Karp-Boss, Boss and Jumars1996; Vetter et al. Reference Vetter, Deming, Jumars and Krieger-Brockett1998; Arnosti, Reference Arnosti2011; Richards & Talbot, Reference Richards and Talbot2013). The key to limiting such losses is containment. Fungi typically manage this through the penetration of solid substrates, whereas osmotrophic prokaryotes exploit the viscosity-dominated fluid dynamics associated with small length scales (<< Re). The diffusive boundary layer (DBL) surrounding micrometre- and sub-micrometre-sized cells is effectively impervious to advective loss, greatly facilitating the rate-limiting steps of both hydrolytic digestion and trans-membrane uptake (Jumars et al. Reference Jumars, Deming, Hill, Karp-Boss, Yager and Dade1993; Langlois et al. Reference Langlois, Andersen, Bohr, Visser, Fishwick and Kiorbøe2009; Arnosti, Reference Arnosti2011). Conversely, the DBL and its facilitation of external digestion and osmotrophy are progressively eroded at larger length scales. Indeed, the primary impediment to osmotrophic feeding in large organisms is not SA:V per se, but the challenge of digesting and incorporating substrate under elevated Re conditions.

The discussion of osmotrophy has been further confused by the term ‘dissolved organic carbon’ (DOC), the substrate on which rangeomorphs are assumed to have fed. Despite early proposals to limit the term to genuinely soluble components (Sharp, Reference Sharp1973), DOC has come to be defined operationally as the reduced carbon content of a filtered water sample, with the pore size of the filter ranging more or less arbitrarily from 0.2 to 1.0 μm (Verdugo et al. Reference Verdugo, Alldredge, Azam, Kirchman, Passow and Santschi2004). As such, DOC now includes a disparate range of materials, from fully dissolved molecules to colloids, gels, viruses and even small microbes.

Further classification of DOC is based on environmental longevity; on the one hand ‘labile DOC’ with residence times of hours to days, and on the other ‘recalcitrant DOC’, which persists from weeks to tens of thousands of years (Hansell, Reference Hansell2013; Follett et al. Reference Follett, Repeta, Rothman, Xu and Santinelli2014). In structural terms, labile DOC is represented by free monomers and small oligomers (<600 daltons) and is the only fraction available for direct osmotrophic uptake. Unsurprisingly, it has limited potential for environmental accumulation, with amino acids and sugars in the modern ocean occurring at concentrations of less than one-billionth of a gram per litre (Hansell, Reference Hansell2013; Moran et al. Reference Moran, Kujawinski, Stubbins, Fatland, Aluwihare, Buchan, Crump, Dorrestein, Dyhrman, Hess, Howe, Longnecker, Medeiros, Niggemann, Obernosterer, Repeta and Waldbauer2016). By contrast, the larger molecules and materials comprising recalcitrant DOC accumulate substantially in the oceans, collectively representing more than 200 times the carbon present in marine biomass. Even so, DOC remains conspicuously dilute in marine environments (from c. 34 to > 80 μmol kg−1), often falling below the threshold necessary to sustain microbially mediated hydrolysis (Arrieta et al. Reference Arrieta, Mayol, Hansman, Herndl, Dittmar and Duarte2015). This alone may contribute to its extended residence time.

There are multiple sinks for recalcitrant DOC in the modern oceans, including both biological remineralization and sedimentary adhesion leading to long-term burial. Of the former, a substantial fraction ends up being captured and consumed by suspension-feeding animals. Benthic sponges, for example, have an extraordinary capacity to extract colloidal DOC from large volumes of water (Yahel et al. Reference Yahel, Sharp, Marie, Häse and Genin2003; de Goeij et al. Reference de Goeij, van Oevelen, Vermeij, Osinga, Middelburg, de Goeij and Admiraal2013; Kahn et al. Reference Kahn, Yahel, Chu, Tunnicliffe and Leys2015), a habit emulated in the pelagic realm by actively swimming salps and tunicates (Flood et al. Reference Flood, Deibel and Morris1992; Sutherland et al. Reference Sutherland, Madin and Stocker2010). None of these DOC feeders can be considered osmotrophic, however, since all of the associated digestion and uptake takes place internally, by endocytic choanocytes in the case of sponges (Leys & Eerkes-Medrano, Reference Leys and Eerkes-Medrano2006) and within a differentiated gut in tunicates and other eumetazoans (e.g. Dishaw et al. Reference Dishaw, Cannon, Litman and Parker2014). At the same time, there is a proportion of DOC that physically aggregates to produce larger particulate organic carbon (POC), with the resulting flakes, gels and transparent exopolymer particles (TEP) available for consumption via conventional eumetazoan-grade capture and ingestion (Camilleri & Ribi, Reference Camilleri and Ribi1986; Verdugo et al. Reference Verdugo, Alldredge, Azam, Kirchman, Passow and Santschi2004; Mari et al. Reference Mari, Passow, Migon, Burd and Legendre2017).

Despite these various non-osmotrophic means of incorporating DOC, it is clear that a range of aquatic eukaryotes do exploit it directly. Labelling experiments, for example, have demonstrated the osmotrophic uptake of acetate, monosaccharides, amino acids and fatty acids by most major invertebrate clades (Wright & Manahan, Reference Wright and Manahan1989; Baines et al. Reference Baines, Fisher and Cole2005; Skikne et al. Reference Skikne, Sherlock and Robison2009; Gori et al. Reference Gori, Grover, Orejas, Sikorski and Ferrier-Pagès2014; Blewett & Goss, Reference Blewett and Goss2017). There are also reports of uncharacterized DOC uptake (e.g. Roditi et al. Reference Roditi, Fisher and Sanudo-Wilhelmy2000; Barnard et al. Reference Barnard, Martineau, Frenette, Dodson and Vincent2006; Rengefors et al. Reference Rengefors, Pålsson, Hansson and Heiberg2008), although it is notable that these substrates were all derived from fresh algal or arthropod lysates. In other words, the osmotrophy observed in modern eukaryotes and metazoans appears to be limited exclusively to the labile, low-molecular-weight DOC that requires no prior digestion; and even then, uptake rates are typically 2–4 orders of magnitude lower than those associated with internalized feeding (Wright & Manahan, Reference Wright and Manahan1989). Reports of aquatic invertebrates feeding osmotrophically on recalcitrant DOC (e.g. Mcmeans et al. Reference Mcmeans, Koussoroplis, Arts and Kainz2015) are likely to involve microbial intermediaries or other means of repackaging leading to internalized digestion (e.g. Camilleri & Ribi, Reference Camilleri and Ribi1986; Höss et al. Reference Höss, Bergtold, Haitzer, Traunspurger and Steinberg2001; Eckert & Pernthaler, Reference Eckert and Pernthaler2014).

Whatever the absolute quality or quantity of food, it only becomes metabolically available once it has been translocated from the external environment into a cell. Although O2, CO2 and a variety of small hydrophobic molecules diffuse more or less freely across phospholipid cell membranes, most organic molecules – including amino acids and monosaccharides – can only be taken up via membrane-embedded transporter proteins. As such, maximum uptake rates are determined by the density and specificity of transporters, and the non-trivial time required for substrate exchange (Confer & Logan, Reference Confer and Logan1991; Karp-Boss et al. Reference Karp-Boss, Boss and Jumars1996). Glucose and amino acid transporters, for example, will saturate under elevated substrate concentrations, limiting the utility of locally enhanced delivery. Even so, the vanishingly low concentrations of free monomers in the modern ocean means that the rate-limiting step for marine osmotrophy will almost always revert to the hydrolytic digestion of more recalcitrant DOC (Confer & Logan, Reference Confer and Logan1991; Moran et al. Reference Moran, Kujawinski, Stubbins, Fatland, Aluwihare, Buchan, Crump, Dorrestein, Dyhrman, Hess, Howe, Longnecker, Medeiros, Niggemann, Obernosterer, Repeta and Waldbauer2016), a process fundamentally at odds with fluid dynamics at macroscopic length scales.

Ultimately, of course, the challenge for all heterotrophic organisms is accessing food. For microbes operating under low Re and Pe regimes, the presence of a thick, essentially permanent DBL means that substrate delivery is dominated by molecular diffusion (Karp-Boss et al. Reference Karp-Boss, Boss and Jumars1996). Despite their extraordinary capacity to digest recalcitrant DOC, the physical inability of osmotrophic microbes to pump or swim substantially through water means they are prone to starvation under oligotrophic conditions (Arrieta et al. Reference Arrieta, Mayol, Hansman, Herndl, Dittmar and Duarte2015). Delivery can be dramatically enhanced in larger organisms simply through the accompanying inertial and turbulent fluid dynamics (Langlois et al. Reference Langlois, Andersen, Bohr, Visser, Fishwick and Kiorbøe2009; Singer et al. Reference Singer, Plotnik and Laflamme2012; Ghisalberti et al. Reference Ghisalberti, Gold, Laflamme, Clapham, Narbonne, Summons, Johnston and Jacobs2014; Guizien & Ghisalberti, Reference Guizien, Ghisalberti, Rossi, Bramanti, Gori and Orejas Saco del Valle2016; Butterfield, Reference Butterfield2018), but at the cost of eroding the stable DBL necessary for external digestion (Vetter et al. Reference Vetter, Deming, Jumars and Krieger-Brockett1998; Arnosti, Reference Arnosti2011; Richards & Talbot, Reference Richards and Talbot2013). In principle, then, macroscopic heterotrophs should be able to make a living under fundamentally more oligotrophic conditions than their microbial counterparts (Ghisalberti et al. Reference Ghisalberti, Gold, Laflamme, Clapham, Narbonne, Summons, Johnston and Jacobs2014), but only if there is an alternative means of digesting recalcitrant food.

2. Osmotrophic rangeomorphs?

Laflamme et al. (Reference Laflamme, Xiao and Kowalewski2009) have argued that Ediacaran rangeomorphs fed osmotrophically via their ‘fractally’ enhanced SA:V properties, noting a marginal overlap with some exceptionally large living prokaryotes. There are problems with this study, however, not least the choice of modern analogues. Of the eight ‘strictly osmotrophic megabacteria’ included in the analysis, only one is actually a free-living osmotroph: the relatively modest-sized (<15 μm diameter) archaeon Staphylothermus marinus. Two others are substantially larger, but known exclusively from the guts of aquatic vertebrates where neither substrate digestion nor advective loss are relevant factors (Sporospirillum praeclarum in tadpoles and Epulopiscium fishelsoni in surgeon fish). None of the remaining five taxa is osmotrophic, or even heterotrophic: Thiomargarita, Achromatium, Beggiatoa and Thiovulum are all sulphur-oxidizing chemoautotrophs, and Prochloron is a cyanobacterial photoautotroph. Indeed, the most likely explanation for the exceptionally large dimensions of these primary producers is the advective delivery of (freely diffusible) CO2 at larger length scales. In any event, there are no extant heterotrophic prokaryotes, free-living or otherwise, that fall within the calculated SA:V range of Ediacaran rangeomorphs (cf. Schulz & Jørgensen, Reference Schulz and Jørgensen2001; Laflamme et al. Reference Laflamme, Xiao and Kowalewski2009, fig. 4b).

The likelihood of osmotrophic feeding in rangeomorphs is also problematic in terms of metabolically available DOC. Although the Proterozoic oceans might well have contained high concentrations of total DOC (e.g. Shields, 2017), any substantial accumulations would have been chemically recalcitrant and unavailable for direct osmotrophic uptake. Arguments for the presence of abundant labile DOC in Proterozoic oceans – due to the absence of metazoan zooplankton and concomitant slow sinking of phytoplankton (Sperling et al. Reference Sperling, Peterson and Laflamme2011) – are incompatible with the voracious consumption of free monomers by heterotrophic microbes, particularly in well-oxygenated surface waters where almost all labile DOC is produced. Labile DOC might well have been intermittently elevated in the vicinity of Ediacaran rangeomorphs (Budd & Jensen, Reference Budd and Jensen2017), but never at the continuously concentrated levels enjoyed by gut-dwelling Sporospirillum and Epulopiscium (cf. Pollak & Montgomery, Reference Pollak and Montgomery1994; Schulz & Jørgensen, Reference Schulz and Jørgensen2001).

The most direct means of resolving the trade-off between effective food delivery (enhanced in larger organisms and elevated Re) and its follow-up incorporation (enhanced at low Re) is to separate the two processes. At a unicellular level, the contained intracellular digestion of POC and DOC by phagocytizing protists by-passes many of the fluid-dynamic challenges faced by osmotrophic prokaryotes, although new ones inevitably arise, not least the elevated Re and turbulence associated with larger eukaryotic cells, and their interference with effective prey capture (Dolan et al. Reference Dolan, Sall, Metcalfe and Gasser2003). At macroscopic length scales, phagocytosis becomes entirely untenable without anatomical or behavioural adaptations for attenuating flow, as seen in the ramifying aquiferous system of sponges (Leys & Eerkes-Medrano, Reference Leys and Eerkes-Medrano2006), the enveloping habit of placozoans (Smith et al. Reference Smith, Pivovarova and Reese2015) or the chambered, often channelized gastrodermal system of cnidarians (Southward, Reference Southward1955; Schick, Reference Schick1991; Harmata et al. Reference Harmata, Parrin, Morrison, McConnell, Bross and Blackstone2013; Raz-Bahat et al. Reference Raz-Bahat, Douek, Moiseeva, Peters and Rinkevich2017; Goldberg, Reference Goldberg, Kloc and Kubiak2018; Steinmetz, Reference Steinmetz2019). Such dynamics presumably account for the exclusively gastrodermal uptake of zooxanthellae in photoendosymbiotic anthozoans. In the case of rangeomorphs, furrows associated with the ‘fractally’ divided integument offer the only potential for comparable levels of isolation on the external surface, although these notably constitute just a fraction of the total surface area.

What the macroscopic size of rangeomorphs certainly does confer is elevated Reynolds and Péclet numbers. Ghisalberti et al. (Reference Ghisalberti, Gold, Laflamme, Clapham, Narbonne, Summons, Johnston and Jacobs2014) presciently recognized the fluid-dynamic implications of large size in Ediacaran macrofossils, noting that the turbulence generated by the interaction of physical currents and an elevated macrobenthos comprehensively overrides any diffusional limits on the delivery of dissolved and suspended resources. Although problematic in terms of osmotrophic feeding, such ‘canopy effects’ are directly applicable to freely diffusible O2 and CO2; indeed, all three of the datasets used by Ghisalberti et al. (Reference Ghisalberti, Gold, Laflamme, Clapham, Narbonne, Summons, Johnston and Jacobs2014, figs 3, S2) to illustrate this principle were specifically measures of oxygen transport. In this light, the most immediate advantage to rangeomorphs adopting a macroscopic habit was access to advective food delivery and gas exchange (Singer et al. Reference Singer, Plotnik and Laflamme2012). Additional ventilatory effects are likely to have been generated by a ciliated epithelium, allowing rangeomorphs to employ their external surface as a breathing device under both high-energy and relatively stagnant physical flow (cf. Short et al. Reference Short, Solari, Ganguly, Powers, Kessler and Goldstein2006; Shapiro et al. Reference Shapiro, Fernandez, Garren, Guasto, Debaillon-Vesque, Kramarsky-Winter, Vardi and Stocker2014; Cavalier-Smith, Reference Cavalier-Smith2017; Dufour & McIlroy, Reference Dufour and McIlroy2017).

2.a. The rangeomorph skeleton

Whatever the particular habits of Ediacaran macrofossils, they were clearly supported by some sort of skeletal superstructure, an anatomical feature that is likely to illuminate other aspects of their biology. In the absence of obvious biomineralization, this has been widely envisaged as a hydrostatic endoskeleton, comparable to the coelomic system of annelid worms (Runnegar, Reference Runnegar1982), the syncytia of giant ‘unicellular’ protists (Seilacher, Reference Seilacher1989, Reference Seilacher1992; Seilacher et al. Reference Seilacher, Grazhdankin and Legouta2003) or a sponge-grade mesenchyme-like mass (Dufour & McIlroy, Reference Dufour and McIlroy2017). More generally, Laflamme et al. (Reference Laflamme, Xiao and Kowalewski2009) have argued that ‘much of the internal body cavity… may have been filled by metabolically inactive material (inorganic, organic, or fluid).’

The key to resolving the nature of the rangeomorph skeleton lies in its taphonomic dissection. Although most rangeomorphs are preserved as more or less 2D bedding-plane imprints – a product of felling, degradational collapse and early diagenetic ‘death-mask’ cementation – there is a notable subset of specimens that have been preserved as conspicuously 3D casts and moulds (Jenkins, Reference Jenkins1985; Fedonkin, Reference Fedonkin and Bengtson1994; Dzik, Reference Dzik2002; Grazhdankin & Seilacher, Reference Grazhdankin and Seilacher2005; Vickers-Rich et al. Reference Vickers-Rich, Yu Ivantsov, Trusler, Narbonne, Hall, Wilson, Greentree, Fedonkin, Elliott, Hoffmann and Schneider2013; Sharp et al. Reference Sharp, Evans, Wilson and Vickers-Rich2017). Such sedimentary infilling points to the presence not only of large internalized chambers, but also chamber walls of sufficient integrity to act as the containing form. In Charnia specimens from the Winter Mountains of the White Sea, for example, substantial parts of the fronds have been infilled with silt early and rapidly enough to capture their full 3D profile (Fedonkin, Reference Fedonkin and Bengtson1994; Grazhdankin, Reference Grazhdankin2004; Dunn et al. Reference Dunn, Wilby, Kenchington, Grazhdankin, Donoghue and Liu2018), even as adjacent unfilled areas collapsed to yield a more typical 2D death-mask (Fig. 2).

Fig. 2. Partially cast 3D specimens of Charnia from the Verkhovka Formation (Winter Mountains, White Sea, Russia), demonstrating the previous presence of water-filled chambers. (a, b) ‘Upper’ and ‘lower’ surfaces of PIN 3993-7018 (with differential transfer of the primary branch casts between the two parts); note that only some parts of this specimen have been infilled with sediment (roughly the left-hand side of (a) and the right-hand side of (b)), with the remainder experiencing a more typical ‘collapse and death-mask’ type of preservation. The three serially repeated lensoid structures preserved on the upper side of the cast (arrows in (a)) potentially represent openings into the chambers; they are not present on the lower ‘fractally’ divided side (b), and are not preserved in collapsed parts of the frond. (c) Cross-section through a silt-cast primary branch of PIN 3993-7018, locally buried in mud and showing anatomical continuity between the chambers and serially repeated lensoid structures (arrowed); line of section indicated by the dotted line in (b). (d) Cross-section through a silt-cast primary branch of PIN 3993-7018, locally buried in cross-laminated silt and showing clear evidence of erosive breaching and loss of the upper body wall; line of section indicated by asterisks in (b). (e) Detail of a further silt-cast, mud-buried specimen (PIN 3992-7020) preserving serially repeated lensoid structures on the upper, non-fractally divided surface (arrows); the full specimen is figured in Fedonkin (Reference Fedonkin and Bengtson1994). Scale bar: (a, b, e) 1 cm; (c) 2.5 mm; and (d) 5 mm. PIN – Palaeontological Institute, Moscow. Photo credits: (a, b, c, e) Dima Grazhdankin; and (d) Alex Liu.

Casting is a common mode of preservation in biomineralized or heavily lignified organisms, but is less expected in ‘soft-bodied’ Ediacaran forms (Seilacher, Reference Seilacher1970; Rex, Reference Rex1985; Retallack, Reference Retallack1994; Maeda et al. Reference Maeda, Kumagae, Matsuoka and Yamazaki2010). Where it does occur, the process will be similarly dependent on chambers with self-supporting walls, but proceeding on fundamentally shorter timescales. Given the rapidly collapsing 2D habit of rangeomorphs in general, it follows that the original contents of the chambers must have been correspondingly fluid. The effectively instantaneous casting of Charnia compartments (Fig. 2) is inconsistent with the original contents having the viscosity of syncytial cytoplasm (cf. Seilacher, Reference Seilacher1989, Reference Seilacher1992), mesenchyme (cf. Dufour & McIlroy, Reference Dufour and McIlroy2017) or coelomic fluids (cf. Runnegar, Reference Runnegar1982), particularly given the rapid wound-repair systems associated with these fully isolated hydrostatic skeletons (e.g. Menzel, Reference Menzel1988; Duckworth, Reference Duckworth2003; Kamran et al. Reference Kamran, Zellner, Kyriazes, Kraus, Reynier and Malamy2017). By far the most likely material filling the compartments of rangeomorphs – and conferring their primary skeletal support – is locally contained low-viscosity seawater.

In the absence of obvious openings in the body wall (Narbonne, Reference Narbonne2004), the route by which such water entered rangeomorph chambers has yet to be identified. One likely possibility is that the conduits were simply too small or ephemeral to fossilize under the associated taphonomic regimes. Even with the fundamentally greater levels of resolution seen in Burgess Shale-type preservation, for example, the ostia and associated aquiferous system of sponges have never been directly preserved (Butterfield, Reference Butterfield2003). The external openings of water-pumping siphonozooids in colonial octocorals can be similarly cryptic, even in living specimens (Fig. 3d) (Hickson, Reference Hickson1883; Brafield, Reference Brafield1969; Nonaka et al. Reference Nonaka, Nakamura, Tsukahara and Reimer2012; Williams et al. Reference Williams, Hoeksema and Van Ofwegen2012), and it is notable that the millimetre-sized siphonozooids of certain pennatulaceans fail to preserve even under the most optimized laboratory burial conditions (Norris, Reference Norris1989).

Fig. 3. Extant anthozoan cnidarians exhibiting features of relevance to the interpretation of Ediacaran rangeomorphs. (a) The modern actiniarian Metridium, demonstrating the disparate range of forms possible by a single specimen depending on the retention and deployment of seawater within the gastrovascular cavity. In the absence of muscle, such an organism would be unable to operate tentacles or a central mouth, although it could still (in principle) function as a suspension-feeding extra-cellular digestion chamber. (b) Scanning electron micrograph (SEM) of the colonial alcyonacean Corallium, showing the surface expression of retracted autozooids (muscle-powered micro-predatory feeding polyps) and cryptically embedded siphonozooids (cilia-powered atentaculate polyps specialized for circulating seawater); the latter are unlikely to be recognizably preserved in the fossil record, even under the most exceptional taphonomic circumstances. (c) Schematic transverse section through a single siphonozooid of the colonial alcyonacean Paragorgia, showing its ciliated water-pumping siphonoglyph (shaded dark blue) and interconnecting gastrovascular canal system (light blue). (d) Schematic longitudinal section of Paragorgia, showing multiple water-pumping siphonozooids with cryptically small external openings (siphonoglyphs shaded dark blue, gastrovascular canals light blue). Scale bar: (a) 2 cm; (b, d) 1 mm; and (c) 0.25 mm. (a) From Batham & Pantin (Reference Batham and Pantin1950), reproduced with permission of the Journal of Experimental Biology. (b) Modified from Nonaka et al. (Reference Nonaka, Nakamura, Tsukahara and Reimer2012). (c, d) Modified from Hickson (Reference Hickson1883).

Considering that such features are only expected under the most exceptional taphonomic circumstances, it is worth revisiting Grazhdankin’s (Reference Grazhdankin2004) original documentation of the 3D Winter Mountains Charnia. Intriguingly, the upwards-facing surfaces of at least two specimens bear serially repeated lensoid structures (Fig. 2a, e, arrows) that are not preserved in corresponding 2D fossils. The continuity of fossil-casting silt through these millimetre-sized structures (Fig. 2c, arrows) points to their likely function as anatomical conduits connecting the internal chambers to surrounding seawater. Taphonomic merger of the ‘upper’ and ‘lower’ surfaces during more typical death-mask preservation readily accounts for their absence in most Charnia fossils (Fig. 4c, d), as well as the misleading impression that the two sides of Charnia fronds were morphologically identical (cf. Dunn et al. Reference Dunn, Wilby, Kenchington, Grazhdankin, Donoghue and Liu2018). Indeed, the preservation of original spatial relationships in these exceptional specimens demonstrates that much of the underlying ‘fractal’ architecture was associated exclusively with the ‘lower’ surface (Fig. 2a, b). In view of their conspicuously mouldic expression, the ‘third-order branches’ of Charnia also appear not to define external morphology, but rather internalized mesentery-like structures – presumably with a primary purpose in expanding internalized surface area (Fig. 4).

Fig. 4. Schematic reconstruction of the constructional and functional anatomy of rangeomorphs, alongside possible taphonomic pathways. (a) Chambered construction with a central mesoglea-like layer (black) supporting a ciliated epithelium; external epidermis (brown) serves as an important locus of high Re/Pe gas exchange, whereas the internalized ‘gastrodermis’ (orange) is optimized for feeding at macroscopic length scales. Overall support is provided by transiently contained seawater (grey). (b) Suspended DOC and POC is cycled through the internalized system via ciliary transport and siphonoglyph-like pumping. Chamber walls are likely to have hosted a diverse, mostly anaerobic microbiome (coloured dots), contributing to gut-like extracellular digestion. (c) Three-dimensional casting of rangeomorph chambers following high-energy erosive breaching of the body wall. (d) Collapse and 2D ‘death-mask’ preservation where the body wall remains intact; the telescoping of spatially separated features onto a single surface yields specimens that appear similar on both surfaces, and obscures key aspects of the original anatomy.

Whether or not the serially repeated lensoid structures represent biological openings in the Charnia integument, it is unlikely that they provided the primary conduit for the infilling sediment. Comparable structures are not recorded in similarly preserved Rangea specimens (e.g. Jenkins, Reference Jenkins1985; Grazhdankin & Seilacher, Reference Grazhdankin and Seilacher2005; Vickers-Rich et al. Reference Vickers-Rich, Yu Ivantsov, Trusler, Narbonne, Hall, Wilson, Greentree, Fedonkin, Elliott, Hoffmann and Schneider2013; Sharp et al. Reference Sharp, Evans, Wilson and Vickers-Rich2017), and such openings in living organisms are almost universally guarded by ciliated and/or contractile cells. Full 3D preservation also requires sufficient ‘draft-through’ flow to deliver the casting sediment prior to degradational collapse (cf. Seilacher, Reference Seilacher1970; Rex, Reference Rex1985; Maeda et al. Reference Maeda, Kumagae, Matsuoka and Yamazaki2010), suggesting substantially larger-scale access to the internalized chambers. Given the high-energy tempestite conditions under which the Winter Mountains Charnia were buried (Grazhdankin, Reference Grazhdankin2004), the most likely route of silt-entraining currents would have been through abrasive breaches in the thin body wall (cf. de Bettignies et al. Reference de Bettignies, Thomsen and Wernberg2012). Indeed, the depositional continuity from fossil-casting silts to overlying cross-laminated horizons in PIN 3993-7018 (Fig. 2d) demonstrates the localized erosion of upwards-facing portions of the frond; it is only where the casts are locally succeeded by low-energy muddy laminae that the differentiated anatomy of this surface is preserved (Figs 2a, e, 4c). In a similar vein, Brasier et al. (Reference Brasier, Liu, Menon, Matthews, McIlroy and Wacey2013) have interpreted features of the high-relief rangeomorphs at Spaniard’s Bay as the consequence of body-wall rupture and/or removal during hydraulic scouring events, and it is notable that the majority of three-dimensionally cast rangeomorphs occur in conspicuously more abrasion- and transport-prone ‘Nama-type’ facies (Grazhdankin, Reference Grazhdankin2004; Grazhdankin & Seilacher, Reference Grazhdankin and Seilacher2005; Vickers-Rich et al. Reference Vickers-Rich, Yu Ivantsov, Trusler, Narbonne, Hall, Wilson, Greentree, Fedonkin, Elliott, Hoffmann and Schneider2013; Sharp et al. Reference Sharp, Evans, Wilson and Vickers-Rich2017).

Rapid sedimentary casting of bag-like compartments is also documented in a range of co-occurring non-rangeomorph Ediacaran taxa. Among the most spectacular examples are in situ, vertically oriented populations of Charniodiscus in Zimnie Gory sections of the White Sea (Grazhdankin, Reference Grazhdankin2014; Ivantsov, Reference Ivantsov2016), reclined but similarly frondose Pambikalbae and Arborea from Nilpena in South Australia (Jenkins & Nedin, 2007; Laflamme et al. Reference Laflamme, Gehling and Droser2018; Dunn et al. Reference Dunn, Liu and Gehling2019; Droser et al. Reference Droser, Tarhan, Evans, Surprenant and Gehling2020) and globular tripartite Ventogyrus from the Onega River area of the White Sea (Ivantsov & Grazhdankin, Reference Ivantsov and Grazhdankin1997; Fedonkin & Ivantsov, Reference Fedonkin and Ivantsov2007). All of these fossils have been variably infilled during event-bed sedimentation, yielding a taphonomic continuum from fully inflated 3D casts through to essentially 2D death-mask imprints. As with rangeomorphs, the form of the casts directly mirrors that of the external moulds, attesting to the thin deformable nature of the chamber walls (Jenkins & Nedin, 2007; Sharp et al. Reference Sharp, Evans, Wilson and Vickers-Rich2017) and the presence of a rapidly displaceable, chamber-filling fluid. Also like rangeomorphs, there is little direct evidence of the openings through which seawater may have entered in life but, similarly, no expectation that these should be recognizably preserved. Again, the infilling sediment was most likely introduced via abrasional breaches in thin body walls. Regardless of any phylogenetic connection to rangeomorphs, the localized 3D casting of these arboreomorphs and other problematica point to a similar grade of chambered construction and hydrostatic support.

2.b. Functional morphology of hydrostatic exoskeletons

The recognition of rangeomorph chambers with direct conduits to the external environment means that the contained seawater was topologically on the outside of the organism, that is, within a bag-like hydrostatic (exo)skeleton. There is nothing particularly exotic about such a system; indeed, it is the primary means of structural support among basal eumetazoans. In actiniarian (sea anemones) and pennatulacean (sea pens, etc.) anthozoans, for example, it is transiently retained seawater in the gastrovascular system that provides the antagonist against which epithelial muscle acts to generate overall form and movement (Fig. 3a) (Batham & Pantin, Reference Batham and Pantin1950; Chapman, Reference Chapman1958). At a more fundamental level, even the muscle can be dispensed with since the flow and containment of seawater within the system is based on embedded ciliary pumps, typically expressed in the form of channelized siphonoglyphs (Fig. 3c, d). This self-supporting bag-like construction offers an exceptionally parsimonious means of assembling a macroscopic organism, not based on costly biominerals, differentiated tissue systems or coelomic fluids, but on environmental water that comes for free. The only substantial costs are a mesoglea-like internal layer that defines the overall form of the inflated chamber (cf. Batham & Pantin, Reference Batham and Pantin1951; Tucker et al. Reference Tucker, Shibata and Blankenship2011) and an enveloping epithelial layer to ensure its integrity (Fig. 4a) (cf. Tyler, Reference Tyler2003; Jonusaite et al. Reference Jonusaite, Donini and Kelly2016). With a charging mechanism based on the plesiomorphic capacity of eukaryotes to pump water (Butterfield, Reference Butterfield2018), such an apparatus provides access to most of the fluid dynamic advantages of large size without the metabolic trade-offs accompanying more complex, carbon-rich body plans (Thingstad et al. Reference Thingstad, Øvreås, Egge, Løvdal and Heldal2005; Acuña et al. Reference Acuña, López-Urrutia and Colin2011; Pitt et al. Reference Pitt, Duarte, Lucas, Sutherland, Condon, Mianzan, Purcell, Robinson and Uye2013).

Most significantly, the middle Ediacaran invention of this bag-like habit solved the problem of conducting extracellular digestion at macroscopic length scales. By containing the process within an essentially impermeable integument, hydrolytic exo-enzymes could now be freely released without advective loss to the environment or competing organisms, even under the turbulent conditions associated with centimetre- to metre-length scales (cf. Sher et al. Reference Sher, Fishman, Melamed-Book, Zhang and Zlotkin2008; Agostini et al. Reference Agostini, Suzuki, Higuchi, Casareto, Yoshinaga, Nakano and Fujimura2012; Raz-Bahat et al. Reference Raz-Bahat, Douek, Moiseeva, Peters and Rinkevich2017; Goldberg, Reference Goldberg, Kloc and Kubiak2018; Steinmetz, Reference Steinmetz2019). Combined with organismal control over the cycling of seawater, the presence of a large-scale holding and mixing vessel provided both the time and hydrodynamic conditions necessary for optimal uptake and digestion, particularly in the presence of substantially expanded mesentery-like surface area (Fig. 4a, b). In modern industrial applications such structures are known as chemical reactors; in biology, they are regularly employed as guts (Penry & Jumars, Reference Penry and Jumars1987).

The most basic type of gut among living animals is the single-opening ‘batch reactor’ of predatory cnidarians, where individual prey items are processed within a (transiently closed) gastrovascular cavity, followed by the regurgitation of undigested remains (Schick, Reference Schick1991; Sher et al. Reference Sher, Fishman, Melamed-Book, Zhang and Zlotkin2008; Schlesinger et al. Reference Schlesinger, Zlotkin, Kramarsky-Winter and Loya2009; Raz-Bahat et al. Reference Raz-Bahat, Douek, Moiseeva, Peters and Rinkevich2017; Steinmetz, Reference Steinmetz2019). Such behaviour, however, is predicated on the availability of suitable prey and a muscle-based means of capturing and manipulating it, for which there is no direct evidence in middle Ediacaran deposits. In this context, the more appropriate model for extracellular digestion is a continuous-flow stirred-tank reactor (CSTR), involving the continuous processing of dissolved or suspended substrate as it passes through a reaction chamber (Penry & Jumars, Reference Penry and Jumars1986). This type of unidirectional water cycling is widely employed by extant cnidarians, where cilia- and siphonoglyph-based pumping is capable of marshalling complex flow paths, even within blind-ended chambers and canals (Southward, Reference Southward1955; Holley & Shelton, Reference Holley and Shelton1984; Schick, Reference Schick1991; Parrin et al. Reference Parrin, Netherton, Bross, McFadden and Blackstone2010; Harmata et al. Reference Harmata, Parrin, Morrison, McConnell, Bross and Blackstone2013; Goldberg, Reference Goldberg, Kloc and Kubiak2018). Fully open-ended unidirectional processing has also been achieved secondarily in colonial pennatulacean and alcyonacean octocorals, through the differentiation and interconnection of water-pumping siphonozooids and stolon systems (Fig. 3c, d) (Hickson, Reference Hickson1883; Brafield, Reference Brafield1969; Williams et al. Reference Williams, Hoeksema and Van Ofwegen2012; Nonaka et al. Reference Nonaka, Nakamura, Tsukahara and Reimer2012), as well as in the atentaculate solitary coral Leptoseris fragilis via the formation of micrometre-scale gastrovascular pores (Schlicter, Reference Schlicter1991). Significantly, active gastrovascular cycling of seawater proceeds even where its skeletal function has been largely superseded by hard skeleton, demonstrating a primary purpose in feeding and internal transport. The ‘fractally’ partitioned hydrostatic exoskeleton of rangeomorphs was similarly suited to such CSTR-like processing.

The Ediacaran introduction of large, gently stirred, semi-enclosed, reaction vessels would have been equally revolutionary from a microbial point of view. Along with the massively expanded area for surface attachment, microbial residence within the rangeomorph chamber system offered both a continuously buffered habitat and essentially unlimited levels of host-delivered resources (Fig. 4b). At the same time, localized containment allowed the direct physiological coupling of otherwise incompatible modes of life. In the bilaterian gut, for example, it is clear that the anaerobic conditions necessary for optimal digestion are maintained both by and for the resident microbiome (Plante, Reference Plante1990; Friedman et al. Reference Friedman, Bittinger, Esipova, Hou, Chau, Jiang, Mesaros, Lund, Liang, FitzGerald, Goulian, Lee, Garcia, Blair, Vinogradov and Wu2018; Litvak et al. Reference Litvak, Byndloss and Bäumler2018), even as the collective ‘holobiome’ takes advantage of a fully oxygenated existence. Comparably steep redox gradients are found in the gastrovascular cavities of extant cnidarians (Agostini et al. Reference Agostini, Suzuki, Higuchi, Casareto, Yoshinaga, Nakano and Fujimura2012), offering similar opportunities for such catabolic partnerships (Viver et al. Reference Viver, Orellana, Hatt, Urdiain, Díaz, Richter, Antón, Avian, Amann, Konstantinidis and Rosselló-Móra2017; Goldberg, Reference Goldberg, Kloc and Kubiak2018). In the case of actiniarian and pennatulacean anthozoans, rhythmic cycling between hypoxic and anoxic conditions within the gastrovascular system (Brafield & Chapman, Reference Brafield and Chapman1967; Chapman, Reference Chapman1972; Jones et al. Reference Jones, Pickthall and Nesbitt1977; Brafield, Reference Brafield1980) reflects the active suppression of oxygen levels, even in the presence of regularly transiting oxygenated seawater (Penry & Jumars, Reference Penry and Jumars1987; Smith & Waltman, Reference Smith and Waltman1995; Agostini et al. Reference Agostini, Suzuki, Higuchi, Casareto, Yoshinaga, Nakano and Fujimura2012). Given the ubiquity of metabolically diverse microbes in the marine realm, the Ediacaran appearance of bag-like rangeomorphs can be viewed as the original evolutionary experiment linking high-Re oxygen-respiring multicellular eukaryotes to a low-Re, hypoxic to anoxic, microbial digester. Such symbioses will have dramatically expanded the capacity of Ediacaran eukaryotes to feed on dilute and/or recalcitrant DOC, while also tapping into the rich physiological, immunological and developmental potential of such redox-sensitive relationships (e.g. McFall-Ngai et al. Reference McFall-Ngai, Hadfield, Bosch, Carey, Domazet-Lošo, Douglas, Dubilier, Eberl, Fukami, Gilbert, Hentschel, King, Kjelleberg, Knoll, Kremer, Mazmanian, Metcalf, Nealson, Pierce, Rawls, Reid, Ruby, Rumpho, Sanders, Tautz and Wernegreen2013; Dishaw et al. Reference Dishaw, Cannon, Litman and Parker2014; Hammarlund, Reference Hammarlund2020).

In addition to optimizing digestion, a large-scale chemical reactor requires reliable delivery of reactants. Although rangeomorphs preserve little direct evidence of their water-processing habits, simply the turbulence generated by the elevated canopy and background currents will have ensured a continuous supply of food and gas exchange (Larsen & RiisgÅrd, Reference Larsen and RiisgÅrd1997; Lassen et al. Reference Lassen, Kortegrd, RiisgÅrd, Friedrichs, Graf and Larsen2006; Singer et al. Reference Singer, Plotnik and Laflamme2012; Ghisalberti et al. Reference Ghisalberti, Gold, Laflamme, Clapham, Narbonne, Summons, Johnston and Jacobs2014). By capitalizing on both the active hydrodynamics of their exposed mostly turbulent outsides, and the unique chemical and microbial milieu of their semi-contained ‘insides’, chamber-forming rangeomorphs invented a fundamentally new way of feeding, breathing and making a living.

2.c. Rangeomorph affiliations

A cnidarian grade of construction does not mean that rangeomorphs were necessarily cnidarians, but it usefully rules out a number of the usual suspects. There is, for example, no feasible means by which the cytoplasmic contents of coenocytic or syncytial eukaryotes could be replaced with sediment on timescales compatible with the 3D preservation of these soft-bodied organisms. Moreover, the absence of any known seaweeds using this sort of chambered, taphonomically castable construction makes an algal interpretation unlikely, even in photic-zone settings. Any convincing case for metazoan affiliation, however, requires the positive identification of diagnostically metazoan features, set in a phylogenetic context. Ignoring problematic ctenophores, there are currently three principal hypotheses for where rangeomorphs might reasonably be positioned within total-group Metazoa: (1) the sister-group of all extant animals (stem-group Metazoa) (Xiao & Laflamme, Reference Xiao and Laflamme2009; Budd & Jensen, Reference Budd and Jensen2017; Dunn et al. Reference Dunn, Liu and Donoghue2017; Darroch et al. Reference Darroch, Smith, Laflamme and Erwin2018); (2) the sister-group of all extant animals minus sponges (stem-group Eumetazoa) (Buss & Seilacher, Reference Buss and Seilacher1994; Dunn et al. Reference Dunn, Liu and Donoghue2017; Hoyal Cuthill & Han, Reference Hoyal Cuthill and Han2018); or (3) the sister-group of all extant cnidarians (stem-group Cnidaria) (Dunn et al. Reference Dunn, Liu and Donoghue2017).

Despite the potential for confusing non-preservation with a true absence of derived characters (Sansom et al. Reference Sansom, Gabbott and Purnell2010), it is clear that rangeomorphs lacked a number of key crown-cnidarian attributes, not least an ability to move or respond usefully to sedimentary inundation. Under comparable levels of event-bed sedimentation, modern actiniarian and pennatulacean cnidarians engage in pronounced whole-organism or whole-colony contraction, an escape response that both fluidizes surrounding sediments and precludes any infilling of gastrovascular compartments (Batham & Pantin, Reference Batham and Pantin1950; Kastendiek, Reference Kastendiek1976; Norris, Reference Norris1989; Holst & Jarms, Reference Holst and Jarms2006; Chimienti et al. Reference Chimienti, Angeletti and Mastrototaro2018). The conspicuously unresponsive habit of rangeomorphs reliably demonstrates their lack of cnidarian-grade muscle. It is also consistent with a lack of (muscle-activated) tentacles and a localized mouth, which in turn implies absence of a cnidarian-grade nerve net, predatory cnidae or predation-based feeding.

In the absence of muscle and associated systems, there appears to be little more to rangeomorphs than perforated bags of water charged by ciliary pumps. But even this represents a fundamental departure from protistan or sponge-grade multicellularity (Arendt et al. Reference Arendt, Benito-Gutierrez, Brunet and Marlow2015). At macroscopic length scales, such a membranous structure can only be realistically achieved with the mechanical reinforcement afforded by specialized intercellular adhesion molecules and a collectivized, extracellular, basement membrane (Tyler, Reference Tyler2003; Nielsen, Reference Nielsen2008; Jonusaite et al. Reference Jonusaite, Donini and Kelly2016). This type of differentiated epithelium is a uniquely eumetazoan feature, and its (inferred) identification in thin-walled rangeomorphs convincingly places these problematic fossils within total-group Eumetazoa (Budd & Jensen, Reference Budd and Jensen2017). The degree to which they can be more precisely resolved depends on the identification of additional phylogenetically informative characters. As pre-muscular, epithelial, tank-based digesters, they offer a compelling model for stem-group eumetazoans. To the extent that macroscopically responsive striated muscle appears to have evolved independently in cnidarians and bilaterians (cf. Steinmetz et al. Reference Steinmetz, Kraus, Larroux, Hammel, Amon-Hassenzahl, Houliston, Wörheide, Nickel, Degnan and Technau2012), they might further be viewed as pre-muscular, pre-predatory, stem-group cnidarians (cf. Marcum & Campbell, Reference Marcum and Campbell1978; Dunn et al. Reference Dunn, Liu and Donoghue2017).

There is much discussion over the nature of the ancestral (eu)metazoan, but the development of a gastrula phase – where the outside surface of a spherical blastula becomes sufficiently invaginated to act as an ‘inside’ – was undoubtedly a key innovation (Nielsen, Reference Nielsen2008; Arendt et al. Reference Arendt, Benito-Gutierrez, Brunet and Marlow2015). Although topologically equivalent to a solitary bag-like cnidarian, neither the gastrula nor its hypothetical ‘gastraea’ counterpart in early metazoan evolution is obviously comparable to macroscopic rangeomorphs, presumably because a large centralized mouth has no function in the absence of muscle, tentacles or, indeed, any food particles large enough to require such an apparatus. In this context, there is a compelling argument for viewing these macroscopic fossils not as single organisms, but as integrated suspension-feeding colonies, broadly analogous to those of extant pennatulacean and alcyonacean octocorals. The serially repeated pore-like structures in three-dimensionally preserved Charnia (Fig. 2) certainly point to colony-like modularity (cf. Dewel, Reference Dewel2000; Dewel et al. Reference Dewel, Dewel and McKinney2001; Hoyal Cuthill & Conway Morris, Reference Hoyal Cuthill and Conway Morris2014; Dececchi et al. Reference Dececchi, Narbonne, Greentree and Laflamme2017; Kenchington et al. Reference Kenchington, Dunn and Wilby2018), while the quantum increase in length scales associated with coloniality would have provided fundamentally enhanced access to water-borne resources without the costs of developing a more sophisticated body plan (cf. Acuña et al. Reference Acuña, López-Urrutia and Colin2011; Pitt et al. Reference Pitt, Duarte, Lucas, Sutherland, Condon, Mianzan, Purcell, Robinson and Uye2013). Unlike predatory octocorals, however, all of the constituent individuals or modules of this hypothetical pre-muscular, colonial rangeomorph would have been deployed as cilia-powered, broadly gastraea-like ‘siphonozooids’, with a primary purpose in circulating water (Figs 3c, d, 4). This does not mean that they are homologous with the siphonozooids of crown-group cnidarians of course (cf. Landing et al. Reference Landing, Antcliffe, Geyer, Kouchinsky, Bowser and Andreas2018). A colonial or modular suspension feeding habit is likely to have evolved independently in any number of stem-group metazoan lineages, just as it has among extant groups (Ryland & Warner, Reference Ryland and Warner1986).

3. Conclusion

Rangeomorphs remain one of the most deeply problematic groups in the fossil record, even as ongoing work reveals novel developmental, anatomical and ecological detail (e.g. Sharp et al. Reference Sharp, Evans, Wilson and Vickers-Rich2017; Kenchington & Wilby, 2017; Dunn et al. Reference Dunn, Wilby, Kenchington, Grazhdankin, Donoghue and Liu2018; Kenchington et al. Reference Kenchington, Dunn and Wilby2018; Liu & Dunn, Reference Liu and Dunn2020). The present study yields yet further levels of biological resolution:

  1. 1. Rangeomorphs were not osmotrophic. The hydrodynamics associated with organisms of this size are physically and biochemically incompatible with such a habit.

  2. 2. Rangeomorphs were supported by a hydrostatic exoskeleton composed of seawater, as demonstrated by the ready castability of internalized chambers during event-bed sedimentation.

  3. 3. The rangeomorph integument was thin-walled, comprising a biomechanically reinforced epithelium and associated mesoglea-like layer. This plastic, bag-like structure was breachable under high-energy siliciclastic sedimentation, but had sufficient integrity to allow three-dimensional casting in silt and sand.

  4. 4. Serially repeated lensoid structures developed (unilaterally) on at least some rangeomorph taxa potentially represent the openings through which seawater circulated in life. Smaller and/or non-preserved channels may also have fulfilled this role, analogous to the cryptic siphonozooids of some modern octocorals.

  5. 5. The flow of seawater through the rangeomorph chamber system is likely to have been driven by collective ciliary pumping, a plesiomorphic property of both metazoans and eukaryotes.

  6. 6. Rangeomorph chambers provided the controlled hydrodynamic and physiological circumstances necessary to conduct extracellular digestion and phagocytosis at macroscopic length scales.

  7. 7. The constructional and functional anatomy of rangeomorphs identifies them as pre-muscular, total-group Eumetazoa.

Prior to the appearance of rangeomorphs there were just two feeding strategies available to free-living heterotrophic organisms: external digestion plus osmotrophy as practiced by prokaryotes and fungi, and the more active capture and internal digestion of phagocytizing protozoans (and sponges). Chamber-forming eumetazoans broke into this ancient duopoly, not by beating microbes at their own game, but through the invention of a revolutionary new technique for harvesting and processing food. By exploiting the unique potential of large size and compartmentalization, eumetazoans tapped into both the turbulent hydrodynamics of their ‘outside’ (discovering an effectively inexhaustible source of both food and gas exchange) and the controlled conditions of their ‘inside’ (allowing both extracellular digestion and ‘osmotrophic’ uptake to be conducted on an industrialized, CSTR-like scale). The key to all of this biological potential was a hydrostatic exoskeleton based on a bag-like epithelium charged by ciliary pumps. In one form or another, such construction underpins the physiology of all eumetazoan life.

Acknowledgements

I thank Graham Budd, Dima Grazhdankin and Alex Liu for valuable discussion, and journal reviewers for their constructive feedback. Fossil images were generously contributed by Dima Grazhdankin, Alex Liu, Jean-Bernard Caron, Phil Wilby and Olga Zhaxybayeva. Figure 3a is reproduced with permission of the Journal of Experimental Biology. This research was supported by Natural Environment Research Council (NERC) grant NE/P002412/1.

References

Acuña, JL, López-Urrutia, Á and Colin, S (2011) Faking giants: the evolution of high prey clearance rates in jellyfishes. Science 333, 1627–9.CrossRefGoogle ScholarPubMed
Agostini, S, Suzuki, Y, Higuchi, T, Casareto, BE, Yoshinaga, K, Nakano, Y and Fujimura, H (2012) Biological and chemical characteristics of the coral gastric cavity. Coral Reefs 31, 147–56.CrossRefGoogle Scholar
Antcliffe, JB and Brasier, MD (2007) Charnia and sea pens are poles apart. Journal of the Geological Society 164, 4951.CrossRefGoogle Scholar
Arendt, D, Benito-Gutierrez, E, Brunet, T and Marlow, H (2015) Gastric pouches and the mucociliary sole: setting the stage for nervous system evolution. Philosophical Transactions of the Royal Society B 370, 20150286.CrossRefGoogle ScholarPubMed
Arnosti, C (2011) Microbial extracellular enzymes and the marine carbon cycle. Annual Review of Marine Science 3, 401–25.CrossRefGoogle ScholarPubMed
Arrieta, JM, Mayol, E, Hansman, RL, Herndl, GJ, Dittmar, T and Duarte, CM (2015) Dilution limits dissolved organic carbon utilization in the deep ocean. Science 348, 331–3.CrossRefGoogle ScholarPubMed
Baines, SB, Fisher, NS and Cole, JJ (2005) Uptake of dissolved organic matter (DOM) and its importance to metabolic requirements of the zebra mussel, Dreissena polymorpha . Limnology and Oceanography 50, 3647.CrossRefGoogle Scholar
Barnard, C, Martineau, C, Frenette, J-J, Dodson, JJ and Vincent, WF (2006) Trophic position of zebra mussel veligers and their use of dissolved organic carbon. Limnology and Oceanography 51, 1473–84.CrossRefGoogle Scholar
Batham, EJ and Pantin, CFA (1950) Muscular and hydrostatic action in the sea-anemone Metridium senile (L.). Journal of Experimental Biology 27, 264–89.CrossRefGoogle Scholar
Batham, EJ and Pantin, CFA (1951) The organization of the muscular system of Metridium senile . Journal of Cell Science 92, 2754.CrossRefGoogle ScholarPubMed
Blewett, TA and Goss, GG (2017) A novel pathway of nutrient absorption in crustaceans: branchial amino acid uptake in the green shore crab (Carcinus maenas). Proceedings of the Royal Society B 284, 20171298.CrossRefGoogle Scholar
Brafield, AE (1969) Water movements in the pennatulid coelenterate Pteroides griseum . Journal of Zoology 158, 317–25.CrossRefGoogle Scholar
Brafield, AE (1980) Oxygen consumption by the sea anemone Calliactis Parasitica (Couch). Journal of Experimental Biology 88, 367–74.CrossRefGoogle Scholar
Brafield, AE and Chapman, G (1967) The respiration of Pteroides griseum (Bohadsch) a pennatulid coelenterate. Journal of Experimental Biology 46, 97104.CrossRefGoogle Scholar
Brasier, MD, Antcliffe, JB and Liu, AG (2012) The architecture of Ediacaran fronds. Palaeontology 55, 1105–24.CrossRefGoogle Scholar
Brasier, MD, Liu, AG, Menon, L, Matthews, JJ, McIlroy, D and Wacey, D (2013) Explaining the exceptional preservation of Ediacaran rangeomorphs from Spaniard’s Bay, Newfoundland: a hydraulic model. Precambrian Research 231, 122–35.CrossRefGoogle Scholar
Budd, GE and Jensen, S (2017) The origin of the animals and a ‘Savannah’ hypothesis for early bilaterian evolution. Biological Reviews 92, 446–73.CrossRefGoogle Scholar
Buss, LW and Seilacher, A (1994) The phylum Vendobionta: a sister group of the Eumetazoa? Paleobiology 20, 14.CrossRefGoogle Scholar
Butterfield, NJ (2003) Exceptional fossil preservation and the Cambrian explosion. Integrative and Comparative Biology 43, 166–77.CrossRefGoogle ScholarPubMed
Butterfield, NJ (2018) Oxygen, animals and aquatic bioturbation: an updated account. Geobiology 16, 316.CrossRefGoogle Scholar
Camilleri, JC and Ribi, G (1986) Leaching of dissolved organic carbon (DOC) from dead leaves, formation of flakes from DOC, and feeding on flakes by crustaceans in mangroves. Marine Biology 91, 337–44.CrossRefGoogle Scholar
Catling, DC, Glein, CR, Zahnle, KJ and McKay, CP (2005) Why O2 is required by complex life on habitable planets and the concept of planetary ‘oxygenation time’. Astrobiology 5, 415–38.CrossRefGoogle Scholar
Cavalier-Smith, T (2017) Origin of animal multicellularity: precursors, causes, consequences—the choanoflagellate/sponge transition, neurogenesis and the Cambrian explosion. Philosophical Transactions of the Royal Society B 372, 20150476.CrossRefGoogle ScholarPubMed
Chapman, G (1958) The hydrostatic skeleton in the invertebrates. Biological Reviews 33, 338–71.CrossRefGoogle Scholar
Chapman, G (1972) A note on the oxygen consumption of Renilla köllikeri, Pfeffer. Comparative Biochemistry and Physiology Part A 42, 63866.Google Scholar
Chimienti, G, Angeletti, L and Mastrototaro, F (2018) Withdrawal behaviour of the red sea pen Pennatula rubra (Cnidaria: Pennatulacea). The European Zoological Journal 85, 6470.CrossRefGoogle Scholar
Confer, DR and Logan, BE (1991) Increased bacterial uptake of macromolecular substrates with fluid shear. Applied and Environmental Microbiology 57, 3093–100.CrossRefGoogle ScholarPubMed
Darroch, SAF, Smith, EF, Laflamme, M and Erwin, DH (2018) Ediacaran extinction and Cambrian explosion. Trends in Ecology & Evolution 33, 653–63.CrossRefGoogle ScholarPubMed
de Bettignies, T, Thomsen, MS and Wernberg, T (2012) Wounded kelps: patterns and susceptibility to breakage. Aquatic Biology 17, 223–33.CrossRefGoogle Scholar
Dececchi, TA, Narbonne, GM, Greentree, C and Laflamme, M (2017) Relating Ediacaran fronds. Paleobiology 43, 171–80.CrossRefGoogle Scholar
de Goeij, JM, van Oevelen, D, Vermeij, MJA, Osinga, R, Middelburg, JJ, de Goeij, AFPM and Admiraal, W (2013) Surviving in a marine desert: the sponge loop retains resources within coral reefs. Science 342, 108–10.CrossRefGoogle Scholar
DeLong, JP, Okie, JG, Moses, ME, Sibly, RM and Brown, JH (2010) Shifts in metabolic scaling, production, and efficiency across major evolutionary transitions of life. Proceedings of the National Academy of Sciences 107, 12941–5.CrossRefGoogle ScholarPubMed
Dewel, RA (2000) Colonial origin for Eumetazoa: major morphological transitions and the origin of bilaterian complexity. Journal of Morphology 243, 3574.3.0.CO;2-#>CrossRefGoogle ScholarPubMed
Dewel, RA, Dewel, WC and McKinney, FK (2001) Diversification of the Metazoa: Ediacarans, colonies, and the origin of eumetazoan complexity by nested modularity. Historical Biology 15, 193218.CrossRefGoogle Scholar
Dishaw, LJ, Cannon, JP, Litman, GW and Parker, W (2014) Immune-directed support of rich microbial communities in the gut has ancient roots. Developmental & Comparative Immunology 47, 3651.CrossRefGoogle ScholarPubMed
Dolan, JR, Sall, N, Metcalfe, A and Gasser, B (2003) Effects of turbulence on the feeding and growth of a marine oligotrich ciliate. Aquatic Microbial Ecology 31, 183–92.CrossRefGoogle Scholar
Droser, ML, Tarhan, LG, Evans, SD, Surprenant, RL and Gehling, JG (2020) Biostratinomy of the Ediacara Member (Rawnsley Quartzite, South Australia): implications for depositional environments, ecology and biology of Ediacara organisms. Interface Focus 10, 20190100.CrossRefGoogle ScholarPubMed
Duckworth, AR (2003) Effect of wound size on the growth and regeneration of two temperate subtidal sponges. Journal of Experimental Marine Biology and Ecology 287, 139–53.CrossRefGoogle Scholar
Dufour, SC and McIlroy, D (2017) Ediacaran pre-placozoan diploblasts in the Avalonian biota: the role of chemosynthesis in the evolution of early animal life. In Earth System Evolution and Early Life: A Celebration of the Work of Martin Brasier (eds AT Brasier, D McIlroy and N McLoughlin), pp. 211–9. Geological Society of London, Special Publication no. 448.CrossRefGoogle Scholar
Dunn, FS, Liu, AG and Donoghue, PCJ (2017) Ediacaran developmental biology. Biological Reviews 93, 914–32.CrossRefGoogle ScholarPubMed
Dunn, FS, Liu, AG and Gehling, JG (2019) Anatomical and ontogenetic reassessment of the Ediacaran frond Arborea arborea and its placement within total group Eumetazoa. Palaeontology 62, 851–65.CrossRefGoogle Scholar
Dunn, FS, Wilby, PR, Kenchington, CG, Grazhdankin, DV, Donoghue, PCJ and Liu, AG (2018) Anatomy of the Ediacaran rangeomorph Charnia masoni . Papers in Palaeontology 5, 157–76.CrossRefGoogle ScholarPubMed
Dzik, J (2002) Possible ctenophoran affinities of the Precambrian “sea-pen” Rangea . Journal of Morphology 252, 315–34.CrossRefGoogle ScholarPubMed
Eckert, EM and Pernthaler, J (2014) Bacterial epibionts of Daphnia: a potential route for the transfer of dissolved organic carbon in freshwater food webs. The ISME Journal 8, 1808–19.CrossRefGoogle ScholarPubMed
Fedonkin, MA (1994) Vendian body fossils and trace fossils. In Early Life on Earth (Nobel Symposium No. 84) (ed. Bengtson, S), pp. 370–88. New York: Columbia University Press.Google Scholar
Fedonkin, MA and Ivantsov, AY (2007) Ventogyrus, a possible siphonophore-like trilobozoan coelenterate from the Vendian sequence (late Neoproterozoic), northern Russia. In The Rise and Fall of the Ediacaran Biota (eds P Vickers-Rich and P Komarower), pp. 187–94. Geological Society of London, Special Publication no. 286.CrossRefGoogle Scholar
Flood, PR, Deibel, D and Morris, CC (1992) Filtration of colloidal melanin from sea water by planktonic tunicates. Nature 355, 630–2.CrossRefGoogle Scholar
Follett, CL, Repeta, DJ, Rothman, DH, Xu, L and Santinelli, C (2014) Hidden cycle of dissolved organic carbon in the deep ocean. Proceedings of the National Academy of Sciences 111, 16706–11.CrossRefGoogle ScholarPubMed
Ford, TD (1958) Pre-Cambrian fossils from Charnwood Forest. Proceedings of the Yorkshire Geological Society 31, 211–7.CrossRefGoogle Scholar
Friedman, ES, Bittinger, K, Esipova, TV, Hou, L, Chau, L, Jiang, J, Mesaros, C, Lund, PJ, Liang, X, FitzGerald, GA, Goulian, M, Lee, D, Garcia, BA, Blair, IA, Vinogradov, SA and Wu, GD (2018) Microbes vs. chemistry in the origin of the anaerobic gut lumen. Proceedings of the National Academy of Sciences 115, 4170–5.CrossRefGoogle ScholarPubMed
Ghisalberti, M, Gold, DA, Laflamme, M, Clapham, ME, Narbonne, GM, Summons, RE, Johnston, DT and Jacobs, DK (2014) Canopy flow analysis reveals the advantage of size in the oldest communities of multicellular eukaryotes. Current Biology 24, 305309.CrossRefGoogle ScholarPubMed
Glazier, DS (2006) The 3/4-power law is not universal: evolution of isometric, ontogenetic metabolic scaling in pelagic animals. BioScience 56, 325–32.CrossRefGoogle Scholar
Goldberg, WM (2018) Coral food, feeding, nutrition, and secretion: a review. In Marine Organisms as Model Systems in Biology and Medicine (eds Kloc, M and Kubiak, JZ), pp. 377421. Chaum, Switzerland: Springer.CrossRefGoogle Scholar
Gori, A, Grover, R, Orejas, C, Sikorski, S and Ferrier-Pagès, C (2014) Uptake of dissolved free amino acids by four cold-water coral species from the Mediterranean Sea. Deep Sea Research Part II 99, 4250.CrossRefGoogle Scholar
Grazhdankin, D (2004) Patterns of distribution in the Ediacaran biotas: facies versus biogeography and evolution. Paleobiology 30, 203–21.2.0.CO;2>CrossRefGoogle Scholar
Grazhdankin, D (2014) Patterns of evolution of the Ediacaran soft-bodied biota. Journal of Paleontology 88, 269–83.CrossRefGoogle Scholar
Grazhdankin, D and Seilacher, A (2005) A re-examination of the Nama-type Vendian organism Rangea schneiderhoehni . Geological Magazine 142, 571–82.CrossRefGoogle Scholar
Guizien, K and Ghisalberti, M (2016) Living in the canopy of the animal forest: physical and biogeochemical aspects. In Marine Animal Forests (eds Rossi, S, Bramanti, L, Gori, A and Orejas Saco del Valle, C), pp. 122. Chaum, Switzerland: Springer International Publishing.Google Scholar
Hammarlund, EU (2020) Harnessing hypoxia as an evolutionary driver of complex multicellularity. Interface Focus 10, 20190101.CrossRefGoogle ScholarPubMed
Hansell, DA (2013) Recalcitrant dissolved organic carbon fractions. Annual Review of Marine Science 5, 421445.CrossRefGoogle ScholarPubMed
Harmata, KL, Parrin, AP, Morrison, PR, McConnell, KK, Bross, LS and Blackstone, NW (2013) Quantitative measures of gastrovascular flow in octocorals and hydroids: toward a comparative biology of transport systems in cnidarians. Invertebrate Biology 132, 291304.CrossRefGoogle Scholar
Hickson, SJ (1883) On the ciliated groove (siphonoglyphe) in the stomodaeum of the alcyonarians. Philosophical Transactions of the Royal Society of London 174, 693705.Google Scholar
Holley, MC and Shelton, GAB (1984) Reversal of the direction of mucus-flow on the ciliated pharynx of a sea anemone. Journal of Experimental Biology 108, 151–61.CrossRefGoogle Scholar
Holst, S and Jarms, G (2006) Responses of solitary and colonial coronate polyps (Cnidaria, Scyphozoa, Coronatae) to sedimentation and burial. Journal of Experimental Marine Biology and Ecology 329, 230–8.CrossRefGoogle Scholar
Höss, S, Bergtold, M, Haitzer, M, Traunspurger, W and Steinberg, CEW (2001) Refractory dissolved organic matter can influence the reproduction of Caenorhabditis elegans (Nematoda) Freshwater Biology 46, 110.CrossRefGoogle Scholar
Hoyal Cuthill, JF and Conway Morris, S (2014) Fractal branching organizations of Ediacaran rangeomorph fronds reveal a lost Proterozoic body plan. Proceedings of the National Academy of Sciences 111, 13122–6.CrossRefGoogle ScholarPubMed
Hoyal Cuthill, JF and Han, J (2018) Cambrian petalonamid Stromatoveris phylogenetically links Ediacaran biota to later animals. Palaeontology 61, 813–23.CrossRefGoogle Scholar
Hurd, CL (2000) Water motion, marine macroalgal physiology, and production. Journal of Phycology 36, 453–72.CrossRefGoogle Scholar
Ivantsov, AY (2016) Reconstruction of Charniodiscus yorgensis (macrobiota from the Vendian of the White Sea). Paleontological Journal 50, 112.CrossRefGoogle Scholar
Ivantsov, AY and Grazhdankin, DV (1997) A new representative of the Petalonamae from the Upper Vendian of the Arkhangelsk region. Paleontological Journal 31, 116.Google Scholar
Jenkins, RJ (1985) The enigmatic Ediacaran (late Precambrian) genus Rangea and related forms. Paleobiology 11, 336–55.CrossRefGoogle Scholar
Jenkins, RJF and Nedin, C (2007) The provenance and palaeobiology of a new multi-vaned, chambered frondose organism from the Ediacaran (later Neoproterozoic) of South Australia. In The Rise and Fall of the Ediacaran Biota (eds Vickers-Rich, P and Komarower, P), pp. 195222. Geological Society of London, Special Publication no. 286.CrossRefGoogle Scholar
Jones, WC, Pickthall, VJ and Nesbitt, SP (1977) A respiratory rhythm in sea anemones. Journal of Experimental Biology 68, 187198.CrossRefGoogle Scholar
Jonusaite, S, Donini, A and Kelly, SP (2016) Occluding junctions of invertebrate epithelia. Journal of Comparative Physiology B 186, 1743.CrossRefGoogle ScholarPubMed
Jumars, PA, Deming, JW, Hill, PS, Karp-Boss, L, Yager, PL and Dade, WB (1993) Physical constraints on marine osmotrophy in an optimal foraging context. Aquatic Microbial Ecology 7, 121–59.Google Scholar
Kahn, AS, Yahel, G, Chu, JWF, Tunnicliffe, V and Leys, SP (2015) Benthic grazing and carbon sequestration by deep-water glass sponge reefs. Limnology and Oceanography 60, 7888.CrossRefGoogle Scholar
Kamran, Z, Zellner, K, Kyriazes, H, Kraus, CM, Reynier, J-B and Malamy, JE (2017) In vivo imaging of epithelial wound healing in the cnidarian Clytia hemisphaerica demonstrates early evolution of purse string and cell crawling closure mechanisms. BMC Developmental Biology 17, 17.CrossRefGoogle ScholarPubMed
Karp-Boss, L, Boss, E and Jumars, PA (1996) Nutrient fluxes to planktonic osmotrophs in the presence of fluid motion. Oceanography and Marine Biology 34, 71108.Google Scholar
Kastendiek, J (1976) Behavior of the sea pansy Renilla kollikeri Pfeffer (Coelenterata: Pennatulacea) and its influence on the distribution and biological interactions of the species. The Biological Bulletin 151, 518–37.CrossRefGoogle Scholar
Kenchington, CG, Dunn, FS and Wilby, PR (2018) Modularity and overcompensatory growth in Ediacaran rangeomorphs demonstrate early adaptations for coping with environmental pressures. Current Biology 28, 3330–6.CrossRefGoogle ScholarPubMed
Kenchington, CG and Wilby, PR (2017) Rangeomorph classification schemes and intra-specific variation: are all characters created equal? In Earth System Evolution and Early Life: A Celebration of the Work of Martin Brasier (eds Brasier, AT, McIlroy, D and McLoughlin, N), pp. 211–50. Geological Society of London, Special Publication no. 448.Google Scholar
Koehl, MAR (1996) When does morphology matter? Annual Review of Ecology and Systematics 27, 501–42.CrossRefGoogle Scholar
Laflamme, M, Gehling, JG and Droser, ML (2018) Deconstructing an Ediacaran frond: three-dimensional preservation of Arborea from Ediacara, South Australia. Journal of Paleontology 92(3), 323–35.CrossRefGoogle Scholar
Laflamme, M, Xiao, S and Kowalewski, M (2009) Osmotrophy in modular Ediacara organisms. Proceedings of the National Academy of Sciences 106, 14438–43.CrossRefGoogle ScholarPubMed
Landing, E, Antcliffe, JB, Geyer, G, Kouchinsky, A, Bowser, SS and Andreas, A (2018) Early evolution of colonial animals (Ediacaran Evolutionary Radiation–Cambrian Evolutionary Radiation–Great Ordovician Biodiversification interval) Earth-Science Reviews 178, 105–35.CrossRefGoogle Scholar
Langlois, V, Andersen, A, Bohr, T, Visser, A, Fishwick, J and Kiorbøe, T (2009) Significance of swimming and feeding currents for nutrient uptake in osmotrophic and interception feeding flagellates. Aquatic Microbial Ecology 54, 3544.CrossRefGoogle Scholar
Larsen, PS and RiisgÅrd, HU (1997) Biomixing generated by benthic filter feeders: a diffusion model for near-bottom phytoplankton depletion. Journal of Sea Research 37, 8190.CrossRefGoogle Scholar
Lassen, J, Kortegrd, M, RiisgÅrd, HU, Friedrichs, M, Graf, G and Larsen, PS (2006) Down-mixing of phytoplankton above filter-feeding mussels—interplay between water flow and biomixing. Marine Ecology Progress Series 314, 7788.CrossRefGoogle Scholar
Leys, SP and Eerkes-Medrano, DI (2006) Feeding in a calcareous sponge: particle uptake by pseudopodia. The Biological Bulletin 211, 157–71.CrossRefGoogle Scholar
Litvak, Y, Byndloss, MX and Bäumler, AJ (2018) Colonocyte metabolism shapes the gut microbiota. Science 362, eaat9076.CrossRefGoogle ScholarPubMed
Liu, AG and Dunn, FS (2020) Filamentous connections between Ediacaran fronds. Current Biology 30, 1322–8.CrossRefGoogle ScholarPubMed
Liu, AG, Kenchington, CG and Mitchell, EG (2015) Remarkable insights into the paleoecology of the Avalonian Ediacaran macrobiota. Gondwana Research 27, 1355–80.CrossRefGoogle Scholar
Maeda, H, Kumagae, T, Matsuoka, H and Yamazaki, Y (2010) Taphonomy of large Canadoceras (ammonoid) shells in the Upper Cretaceous Series in South Sakhalin, Russia. Paleontological Research 14, 5668.CrossRefGoogle Scholar
Marcum, BA and Campbell, RD (1978) Development of Hydra lacking nerve and interstitial cells. Journal of Cell Science 29, 1733.CrossRefGoogle ScholarPubMed
Mari, X, Passow, U, Migon, C, Burd, AB and Legendre, L (2017) Transparent exopolymer particles: effects on carbon cycling in the ocean. Progress in Oceanography 151, 1337.CrossRefGoogle Scholar
McFall-Ngai, M, Hadfield, MG, Bosch, TCG, Carey, HV, Domazet-Lošo, T, Douglas, AE, Dubilier, N, Eberl, G, Fukami, T, Gilbert, SF, Hentschel, U, King, N, Kjelleberg, S, Knoll, AH, Kremer, N, Mazmanian, SK, Metcalf, JL, Nealson, K, Pierce, NE, Rawls, JF, Reid, A, Ruby, EG, Rumpho, M, Sanders, JG, Tautz, D and Wernegreen, JJ (2013) Animals in a bacterial world, a new imperative for the life sciences. Proceedings of the National Academy of Sciences 110, 3229–36.CrossRefGoogle Scholar
Mcmeans, BC, Koussoroplis, A-M, Arts, MT and Kainz, MJ (2015) Terrestrial dissolved organic matter supports growth and reproduction of Daphnia magna when algae are limiting. Journal of Plankton Research 37, 1201–09.Google Scholar
McMenamin, M (1993) Osmotrophy in fossil protoctists and early animals. Invertebrate Reproduction & Development 23, 165–6.CrossRefGoogle Scholar
Menzel, D (1988) How do giant plant cells cope with injury?—The wound response in siphonous green algae. Protoplasma 144, 7391.CrossRefGoogle Scholar
Moran, MA, Kujawinski, EB, Stubbins, A, Fatland, R, Aluwihare, LI, Buchan, A, Crump, BC, Dorrestein, PC, Dyhrman, ST, Hess, NJ, Howe, B, Longnecker, K, Medeiros, PM, Niggemann, J, Obernosterer, I, Repeta, DJ and Waldbauer, JR (2016) Deciphering ocean carbon in a changing world. Proceedings of the National Academy of Sciences 113, 3143–51.CrossRefGoogle Scholar
Narbonne, GM (2004) Modular construction of early Ediacaran complex life forms. Science 305, 1141–44.CrossRefGoogle ScholarPubMed
Nielsen, C (2008) Six major steps in animal evolution: are we derived sponge larvae? Evolution & Development 10, 241–57.CrossRefGoogle ScholarPubMed
Nonaka, M, Nakamura, M, Tsukahara, M and Reimer, JD (2012) Histological examination of precious corals from the Ryukyu Archipelago. Journal of Marine Biology 2012, 114.CrossRefGoogle Scholar
Norris, RD (1989) Cnidarian taphonomy and affinities of the Ediacara biota. Lethaia 22, 381–93.CrossRefGoogle Scholar
Parrin, AP, Netherton, SE, Bross, LS, McFadden, CS and Blackstone, NW (2010) Circulation of fluids in the gastrovascular system of a stoloniferan octocoral. Biological Bulletin 219, 112–21.CrossRefGoogle ScholarPubMed
Penry, DL and Jumars, PA (1986) Chemical reactor analysis and optimal digestion. BioScience 36, 310–15.CrossRefGoogle Scholar
Penry, DL and Jumars, PA (1987) Modeling animal guts as chemical reactors. The American Naturalist 129, 6996.CrossRefGoogle Scholar
Pitt, KA, Duarte, CM, Lucas, CH, Sutherland, KR, Condon, RH, Mianzan, H, Purcell, JE, Robinson, KL and Uye, S-I (2013) Jellyfish body plans provide allometric advantages beyond low carbon content. PLoS ONE 8, e72683.CrossRefGoogle ScholarPubMed
Plante, CJ (1990) Digestive associations between marine detritivores and bacteria. Annual Review of Ecology and Systematics 21, 93127.CrossRefGoogle Scholar
Pollak, PE and Montgomery, WL (1994) Giant bacterium (Epulopiscium fishelsoni) influences digestive enzyme activity of an herbivorous surgeonfish (Acanthurus nigrofuscus). Comparative Biochemistry and Physiology Part A 108, 657–62.CrossRefGoogle Scholar
Raz-Bahat, M, Douek, J, Moiseeva, E, Peters, EC and Rinkevich, B (2017) The digestive system of the stony coral Stylophora pistillata . Cell and Tissue Research 368, 311–23.CrossRefGoogle ScholarPubMed
Rengefors, K, Pålsson, C, Hansson, L and Heiberg, L (2008) Cell lysis of competitors and osmotrophy enhance growth of the bloom-forming alga Gonyostomum semen . Aquatic Microbial Ecology 51, 8796.CrossRefGoogle Scholar
Retallack, GJ (1994) Were the Ediacaran fossils lichens? Paleobiology 20, 523–44.CrossRefGoogle Scholar
Rex, G (1985) A laboratory flume investigation of the formation of fossil stem fills. Sedimentology 32, 245–55.CrossRefGoogle Scholar
Richards, TA and Talbot, NJ (2013) Horizontal gene transfer in osmotrophs: playing with public goods. Nature Reviews Microbiology 11, 720–7.CrossRefGoogle ScholarPubMed
Roditi, HA, Fisher, NS and Sanudo-Wilhelmy, SA (2000) Uptake of dissolved organic carbon and trace elements by zebra mussels. Nature 407, 7880.CrossRefGoogle ScholarPubMed
Runnegar, B (1982) Oxygen requirements, biology and phylogenetic significance of the late Precambrian worm Dickinsonia, and the evolution of the burrowing habit. Alcheringa 6, 223–39.CrossRefGoogle Scholar
Ryland, JS and Warner, GF (1986) Growth and form in modular animals: ideas on the size and arrangement of zooids. Philosophical Transactions of the Royal Society of London B 313, 5376.Google Scholar
Sansom, RS, Gabbott, SE and Purnell, MA (2010) Non-random decay of chordate characters causes bias in fossil interpretation. Nature 463, 797800.CrossRefGoogle ScholarPubMed
Schick, MJ (1991) A Functional Biology of Sea Anemones. London: Chapman & Hall, 395 pp.CrossRefGoogle Scholar
Schlesinger, A, Zlotkin, E, Kramarsky-Winter, E and Loya, Y (2009) Cnidarian internal stinging mechanism. Proceedings of the Royal Society of London B 276, 1063–7.Google ScholarPubMed
Schlicter, D (1991) A perforated gastrovascular cavity in the symbiotic deep-water coral Leptoseris fragilis: a new strategy to optimize heterotrophic nutrition. Helgolander Meeresuntersuchungen 45, 423–43.CrossRefGoogle Scholar
Schulz, HN and Jørgensen, BB (2001) Big bacteria. Annual Reviews in Microbiology 55, 105–37.CrossRefGoogle ScholarPubMed
Sebens, KP (1987) The ecology of indeterminate growth in animals. Annual Review of Ecology and Systematics 18, 371407.CrossRefGoogle Scholar
Seilacher, A (1970) Preservational history of ceratite shells. Palaeontology 14, 1621.Google Scholar
Seilacher, A (1989) Vendozoa: organismic construction in the Proterozoic biosphere. Lethaia 22, 229–39.CrossRefGoogle Scholar
Seilacher, A (1992) Vendobionta and Psammocorallia: lost constructions of Precambrian evolution. Journal of the Geological Society 149, 607–13.CrossRefGoogle Scholar
Seilacher, A, Grazhdankin, D and Legouta, A (2003) Ediacaran biota: the dawn of animal life in the shadow of giant protists. Paleontological Research 7, 4354.CrossRefGoogle Scholar
Shapiro, OH, Fernandez, VI, Garren, M, Guasto, JS, Debaillon-Vesque, FP, Kramarsky-Winter, E, Vardi, A and Stocker, R (2014) Vortical ciliary flows actively enhance mass transport in reef corals. Proceedings of the National Academy of Sciences 111, 13391–6.CrossRefGoogle ScholarPubMed
Sharp, AC, Evans, AR, Wilson, SA and Vickers-Rich, P (2017) First non-destructive internal imaging of Rangea, an icon of complex Ediacaran life. Precambrian Research 299, 303–08.CrossRefGoogle Scholar
Sharp, JH (1973) Size classes of organic carbon in seawater. Limnology and Oceanography 18, 441–7.CrossRefGoogle Scholar
Sher, D, Fishman, Y, Melamed-Book, N, Zhang, M and Zlotkin, E (2008) Osmotically driven prey disintegration in the gastrovascular cavity of the green hydra by a pore-forming protein. FASEB Journal 22, 207–14.CrossRefGoogle ScholarPubMed
Shields, GA (2017) Earth system transition during the Tonian–Cambrian interval of biological innovation: nutrients, climate, oxygen and the marine organic carbon capacitor. In Earth System Evolution and Early Life: A Celebration of the Work of Martin Brasier (eds Brasier, AT, McIlroy, D and McLoughlin, N), pp. 167–77. Geological Society of London, Special Publication no. 448.CrossRefGoogle Scholar
Short, MB, Solari, CA, Ganguly, S, Powers, TR, Kessler, JO and Goldstein, RE (2006) Flows driven by flagella of multicellular organisms enhance long-range molecular transport. Proceedings of the National Academy of Sciences 103, 8315–9.CrossRefGoogle ScholarPubMed
Singer, A, Plotnik, R, and Laflamme, M (2012) Experimental fluid dynamics of an Ediacaran frond. Palaeontologia Electronica 15, 114.Google Scholar
Skikne, SA, Sherlock, RE and Robison, BH (2009) Uptake of dissolved organic matter by ephyrae of two species of scyphomedusae. Journal of Plankton Research 31, 1563–70.CrossRefGoogle Scholar
Smith, CL, Pivovarova, N and Reese, TS (2015) Coordinated feeding behavior in Trichoplax, an animal without synapses. PLoS ONE 10, e0136098.CrossRefGoogle ScholarPubMed
Smith, HL and Waltman, PE (1995) The Theory of the Chemostat: Dynamics of Microbial Competition. Cambridge: Cambridge University Press, 332 pp.CrossRefGoogle Scholar
Solari, CA, Kessler, JO and Goldstein, RE (2007) Motility, mixing, and multicellularity. Genetic Programming and Evolvable Machines 8, 115–29.CrossRefGoogle Scholar
Southward, AJ (1955) Observations on the ciliary currents of the jelly-fish Aurelia aurita L. Journal of the Marine Biological Association of the United Kingdom 34, 201–16.CrossRefGoogle Scholar
Sperling, EA, Peterson, KJ and Laflamme, M (2011) Rangeomorphs, Thectardis (Porifera?) and dissolved organic carbon in the Ediacaran oceans. Geobiology 9, 2433.CrossRefGoogle ScholarPubMed
Steinmetz, PRH (2019) A non-bilaterian perspective on the development and evolution of animal digestive systems. Cell and Tissue Research 377, 321–39.CrossRefGoogle ScholarPubMed
Steinmetz, PRH, Kraus, JEM, Larroux, C, Hammel, JU, Amon-Hassenzahl, A, Houliston, E, Wörheide, G, Nickel, M, Degnan, BM and Technau, U (2012) Independent evolution of striated muscles in cnidarians and bilaterians. Nature 487, 231–4.CrossRefGoogle ScholarPubMed
Sutherland, KR, Madin, LP and Stocker, R (2010) Filtration of submicrometer particles by pelagic tunicates. Proceedings of the National Academy of Sciences 107, 15129–34.CrossRefGoogle ScholarPubMed
Thingstad, TF, Øvreås, L, Egge, JK, Løvdal, T and Heldal, M (2005) Use of non-limiting substrates to increase size; a generic strategy to simultaneously optimize uptake and minimize predation in pelagic osmotrophs? Ecology Letters 8, 675–82.CrossRefGoogle Scholar
Tucker, RP, Shibata, B and Blankenship, TN (2011) Ultrastructure of the mesoglea of the sea anemone Nematostella vectensis (Edwardsiidae). Invertebrate Biology 130, 1124.CrossRefGoogle Scholar
Tyler, S (2003) Epithelium—the primary building block for metazoan complexity. Integrative and Comparative Biology 43, 5563.CrossRefGoogle ScholarPubMed
Verdugo, P, Alldredge, AL, Azam, F, Kirchman, DL, Passow, U and Santschi, PH (2004) The oceanic gel phase: a bridge in the DOM–POM continuum. Marine Chemistry 92, 6785.CrossRefGoogle Scholar
Vetter, YA, Deming, JW, Jumars, PA and Krieger-Brockett, BB (1998) A predictive model of bacterial foraging by means of freely released extracellular enzymes. Microbial Ecology 36, 7592.CrossRefGoogle ScholarPubMed
Vickers-Rich, P, Yu Ivantsov, A, Trusler, PW, Narbonne, GM, Hall, M, Wilson, SA, Greentree, C, Fedonkin, MA, Elliott, DA, Hoffmann, KH and Schneider, GIC (2013) Reconstructing Rangea: new discoveries from the Ediacaran of southern Namibia. Journal of Paleontology 87, 115.CrossRefGoogle Scholar
Viver, T, Orellana, LH, Hatt, JK, Urdiain, M, Díaz, S, Richter, M, Antón, J, Avian, M, Amann, R, Konstantinidis, KT and Rosselló-Móra, R (2017) The low diverse gastric microbiome of the jellyfish Cotylorhiza tuberculata is dominated by four novel taxa. Environmental Microbiology 19, 3039–58.CrossRefGoogle ScholarPubMed
Williams, GC, Hoeksema, BW and Van Ofwegen, LP (2012) A fifth morphological polyp in pennatulacean octocorals, with a review of polyp polymorphism in the genera Pennatula and Pteroeides (Anthozoa: Pennatulidae). Zoological Studies 51, 1006–17.Google Scholar
Wood, DA, Dalrymple, RW, Narbonne, GM, Gehling, JG and Clapham, ME (2003) Paleoenvironmental analysis of the late Neoproterozoic Mistaken Point and Trepassey formations, southeastern Newfoundland. Canadian Journal of Earth Sciences 40, 1375–91.CrossRefGoogle Scholar
Wright, SH and Manahan, DT (1989) Integumental nutrient uptake by aquatic organisms. Annual Review of Physiology 51, 585600.CrossRefGoogle ScholarPubMed
Xiao, S and Laflamme, M (2009) On the eve of animal radiation: phylogeny, ecology and evolution of the Ediacara biota. Trends in Ecology & Evolution 24, 3140.CrossRefGoogle ScholarPubMed
Yahel, G, Sharp, JH, Marie, D, Häse, C and Genin, A (2003) In situ feeding and element removal in the symbiont-bearing sponge Theonella swinhoei: bulk DOC is the major source for carbon. Limnology and Oceanography 48, 141–9.CrossRefGoogle Scholar
Figure 0

Fig. 1. Rangeomorph taxa illustrating the characteristic fractal-like branching and diversity of overall form. (a) Charnia masoni, type specimen, from Charnwood Forest, UK. (b) Rangea schneiderhoehni, type specimen, from Namibia. (c) Hapsidophyllas flexibilis, from SE Newfoundland. (d) Fractofusus misrae from SE Newfoundland. (e) Bradgatia sp. from SE Newfoundland. Scale bar: (a, e) 2 cm; (b) 1.5 cm; (c) 4 cm; and (d) 3 cm. Photo credits: (a) Phil Wilby; (b) Dima Grazhdankin; (c) Olga Zhaxybayeva; (d) Alex Liu; and (e) Jean-Bernard Caron.

Figure 1

Fig. 2. Partially cast 3D specimens of Charnia from the Verkhovka Formation (Winter Mountains, White Sea, Russia), demonstrating the previous presence of water-filled chambers. (a, b) ‘Upper’ and ‘lower’ surfaces of PIN 3993-7018 (with differential transfer of the primary branch casts between the two parts); note that only some parts of this specimen have been infilled with sediment (roughly the left-hand side of (a) and the right-hand side of (b)), with the remainder experiencing a more typical ‘collapse and death-mask’ type of preservation. The three serially repeated lensoid structures preserved on the upper side of the cast (arrows in (a)) potentially represent openings into the chambers; they are not present on the lower ‘fractally’ divided side (b), and are not preserved in collapsed parts of the frond. (c) Cross-section through a silt-cast primary branch of PIN 3993-7018, locally buried in mud and showing anatomical continuity between the chambers and serially repeated lensoid structures (arrowed); line of section indicated by the dotted line in (b). (d) Cross-section through a silt-cast primary branch of PIN 3993-7018, locally buried in cross-laminated silt and showing clear evidence of erosive breaching and loss of the upper body wall; line of section indicated by asterisks in (b). (e) Detail of a further silt-cast, mud-buried specimen (PIN 3992-7020) preserving serially repeated lensoid structures on the upper, non-fractally divided surface (arrows); the full specimen is figured in Fedonkin (1994). Scale bar: (a, b, e) 1 cm; (c) 2.5 mm; and (d) 5 mm. PIN – Palaeontological Institute, Moscow. Photo credits: (a, b, c, e) Dima Grazhdankin; and (d) Alex Liu.

Figure 2

Fig. 3. Extant anthozoan cnidarians exhibiting features of relevance to the interpretation of Ediacaran rangeomorphs. (a) The modern actiniarian Metridium, demonstrating the disparate range of forms possible by a single specimen depending on the retention and deployment of seawater within the gastrovascular cavity. In the absence of muscle, such an organism would be unable to operate tentacles or a central mouth, although it could still (in principle) function as a suspension-feeding extra-cellular digestion chamber. (b) Scanning electron micrograph (SEM) of the colonial alcyonacean Corallium, showing the surface expression of retracted autozooids (muscle-powered micro-predatory feeding polyps) and cryptically embedded siphonozooids (cilia-powered atentaculate polyps specialized for circulating seawater); the latter are unlikely to be recognizably preserved in the fossil record, even under the most exceptional taphonomic circumstances. (c) Schematic transverse section through a single siphonozooid of the colonial alcyonacean Paragorgia, showing its ciliated water-pumping siphonoglyph (shaded dark blue) and interconnecting gastrovascular canal system (light blue). (d) Schematic longitudinal section of Paragorgia, showing multiple water-pumping siphonozooids with cryptically small external openings (siphonoglyphs shaded dark blue, gastrovascular canals light blue). Scale bar: (a) 2 cm; (b, d) 1 mm; and (c) 0.25 mm. (a) From Batham & Pantin (1950), reproduced with permission of the Journal of Experimental Biology. (b) Modified from Nonaka et al. (2012). (c, d) Modified from Hickson (1883).

Figure 3

Fig. 4. Schematic reconstruction of the constructional and functional anatomy of rangeomorphs, alongside possible taphonomic pathways. (a) Chambered construction with a central mesoglea-like layer (black) supporting a ciliated epithelium; external epidermis (brown) serves as an important locus of high Re/Pe gas exchange, whereas the internalized ‘gastrodermis’ (orange) is optimized for feeding at macroscopic length scales. Overall support is provided by transiently contained seawater (grey). (b) Suspended DOC and POC is cycled through the internalized system via ciliary transport and siphonoglyph-like pumping. Chamber walls are likely to have hosted a diverse, mostly anaerobic microbiome (coloured dots), contributing to gut-like extracellular digestion. (c) Three-dimensional casting of rangeomorph chambers following high-energy erosive breaching of the body wall. (d) Collapse and 2D ‘death-mask’ preservation where the body wall remains intact; the telescoping of spatially separated features onto a single surface yields specimens that appear similar on both surfaces, and obscures key aspects of the original anatomy.